/[MITgcm]/manual/s_overview/text/manual.tex
ViewVC logotype

Contents of /manual/s_overview/text/manual.tex

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


Revision 1.3 - (show annotations) (download) (as text)
Wed Oct 10 16:48:45 2001 UTC (23 years, 9 months ago) by cnh
Branch: MAIN
Changes since 1.2: +80 -14 lines
File MIME type: application/x-tex
Updates to part1 for figures

1 % $Header: /u/u0/gcmpack/mitgcmdoc/part1/manual.tex,v 1.2 2001/10/09 10:48:03 cnh Exp $
2 % $Name: $
3 %\usepackage{oldgerm}
4 % I commented the following because it introduced excessive white space
5 %\usepackage{palatcm} % better PDF
6 % page headers and footers
7 %\pagestyle{fancy}
8 % referencing
9 %% \newcommand{\refequ}[1]{equation (\ref{equ:#1})}
10 %% \newcommand{\refequbig}[1]{Equation (\ref{equ:#1})}
11 %% \newcommand{\reftab}[1]{Tab.~\ref{tab:#1}}
12 %% \newcommand{\reftabno}[1]{\ref{tab:#1}}
13 %% \newcommand{\reffig}[1]{Fig.~\ref{fig:#1}}
14 %% \newcommand{\reffigno}[1]{\ref{fig:#1}}
15 % stuff for psfrag
16 %% \newcommand{\textinfigure}[1]{{\footnotesize\textbf{\textsf{#1}}}}
17 %% \newcommand{\mathinfigure}[1]{\small\ensuremath{{#1}}}
18 % This allows numbering of subsubsections
19 % This changes the the chapter title
20 %\renewcommand{\chaptername}{Section}
21
22
23 %%%% \documentclass[12pt]{book}
24 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
25 %%%% \usepackage{amsmath}
26 %%%% \usepackage{html}
27 %%%% \usepackage{epsfig}
28 %%%% \usepackage{graphics,subfigure}
29 %%%% \usepackage{array}
30 %%%% \usepackage{multirow}
31 %%%% \usepackage{fancyhdr}
32 %%%% \usepackage{psfrag}
33
34 %%%% %TCIDATA{OutputFilter=Latex.dll}
35 %%%% %TCIDATA{LastRevised=Thursday, October 04, 2001 14:41:22}
36 %%%% %TCIDATA{<META NAME="GraphicsSave" CONTENT="32">}
37 %%%% %TCIDATA{Language=American English}
38
39 %%%% \fancyhead{}
40 %%%% \fancyhead[LO]{\slshape \rightmark}
41 %%%% \fancyhead[RE]{\slshape \leftmark}
42 %%%% \fancyhead[RO,LE]{\thepage}
43 %%%% \fancyfoot[CO,CE]{\today}
44 %%%% \fancyfoot[RO,LE]{ }
45 %%%% \renewcommand{\headrulewidth}{0.4pt}
46 %%%% \renewcommand{\footrulewidth}{0.4pt}
47 %%%% \setcounter{secnumdepth}{3}
48 %%%% \input{tcilatex}
49 %%%%
50 %%%% \begin{document}
51 %%%%
52 %%%% \tableofcontents
53 %%%%
54 %%%% \pagebreak
55
56 %%%% \part{MIT GCM basics}
57
58 % Section: Overview
59
60 % $Header: /u/u0/gcmpack/mitgcmdoc/part1/manual.tex,v 1.2 2001/10/09 10:48:03 cnh Exp $
61 % $Name: $
62
63 \section{Introduction}
64
65 This documentation provides the reader with the information necessary to
66 carry out numerical experiments using MITgcm. It gives a comprehensive
67 description of the continuous equations on which the model is based, the
68 numerical algorithms the model employs and a description of the associated
69 program code. Along with the hydrodynamical kernel, physical and
70 biogeochemical parameterizations of key atmospheric and oceanic processes
71 are available. A number of examples illustrating the use of the model in
72 both process and general circulation studies of the atmosphere and ocean are
73 also presented.
74
75 MITgcm has a number of novel aspects:
76
77 \begin{itemize}
78 \item it can be used to study both atmospheric and oceanic phenomena; one
79 hydrodynamical kernel is used to drive forward both atmospheric and oceanic
80 models - see fig%
81 \marginpar{
82 Fig.1 One model}\ref{fig:onemodel}
83
84 %% CNHbegin
85 \input{part1/one_model_figure}
86 %% CNHend
87
88 \item it has a non-hydrostatic capability and so can be used to study both
89 small-scale and large scale processes - see fig %
90 \marginpar{
91 Fig.2 All scales}\ref{fig:all-scales}
92
93 %% CNHbegin
94 \input{part1/all_scales_figure}
95 %% CNHend
96
97 \item finite volume techniques are employed yielding an intuitive
98 discretization and support for the treatment of irregular geometries using
99 orthogonal curvilinear grids and shaved cells - see fig %
100 \marginpar{
101 Fig.3 Finite volumes}\ref{fig:finite-volumes}
102
103 %% CNHbegin
104 \input{part1/fvol_figure}
105 %% CNHend
106
107 \item tangent linear and adjoint counterparts are automatically maintained
108 along with the forward model, permitting sensitivity and optimization
109 studies.
110
111 \item the model is developed to perform efficiently on a wide variety of
112 computational platforms.
113 \end{itemize}
114
115 Key publications reporting on and charting the development of the model are
116 listed in an Appendix.
117
118 We begin by briefly showing some of the results of the model in action to
119 give a feel for the wide range of problems that can be addressed using it.
120 \pagebreak
121
122 % $Header: /u/u0/gcmpack/mitgcmdoc/part1/manual.tex,v 1.2 2001/10/09 10:48:03 cnh Exp $
123 % $Name: $
124
125 \section{Illustrations of the model in action}
126
127 The MITgcm has been designed and used to model a wide range of phenomena,
128 from convection on the scale of meters in the ocean to the global pattern of
129 atmospheric winds - see fig.2\ref{fig:all-scales}. To give a flavor of the
130 kinds of problems the model has been used to study, we briefly describe some
131 of them here. A more detailed description of the underlying formulation,
132 numerical algorithm and implementation that lie behind these calculations is
133 given later. Indeed many of the illustrative examples shown below can be
134 easily reproduced: simply download the model (the minimum you need is a PC
135 running linux, together with a FORTRAN\ 77 compiler) and follow the examples
136 described in detail in the documentation.
137
138 \subsection{Global atmosphere: `Held-Suarez' benchmark}
139
140 A novel feature of MITgcm is its ability to simulate both atmospheric and
141 oceanographic flows at both small and large scales.
142
143 Fig.E1a.\ref{fig:eddy_cs} shows an instantaneous plot of the 500$mb$
144 temperature field obtained using the atmospheric isomorph of MITgcm run at
145 2.8$^{\circ }$ resolution on the cubed sphere. We see cold air over the pole
146 (blue) and warm air along an equatorial band (red). Fully developed
147 baroclinic eddies spawned in the northern hemisphere storm track are
148 evident. There are no mountains or land-sea contrast in this calculation,
149 but you can easily put them in. The model is driven by relaxation to a
150 radiative-convective equilibrium profile, following the description set out
151 in Held and Suarez; 1994 designed to test atmospheric hydrodynamical cores -
152 there are no mountains or land-sea contrast.
153
154 %% CNHbegin
155 \input{part1/cubic_eddies_figure}
156 %% CNHend
157
158 As described in Adcroft (2001), a `cubed sphere' is used to discretize the
159 globe permitting a uniform gridding and obviated the need to fourier filter.
160 The `vector-invariant' form of MITgcm supports any orthogonal curvilinear
161 grid, of which the cubed sphere is just one of many choices.
162
163 Fig.E1b shows the 5-year mean, zonally averaged potential temperature, zonal
164 wind and meridional overturning streamfunction from a 20-level version of
165 the model. It compares favorable with more conventional spatial
166 discretization approaches.
167
168 A regular spherical lat-lon grid can also be used.
169
170 %% CNHbegin
171 \input{part1/hs_zave_u_figure}
172 %% CNHend
173
174 \subsection{Ocean gyres}
175
176 Baroclinic instability is a ubiquitous process in the ocean, as well as the
177 atmosphere. Ocean eddies play an important role in modifying the
178 hydrographic structure and current systems of the oceans. Coarse resolution
179 models of the oceans cannot resolve the eddy field and yield rather broad,
180 diffusive patterns of ocean currents. But if the resolution of our models is
181 increased until the baroclinic instability process is resolved, numerical
182 solutions of a different and much more realistic kind, can be obtained.
183
184 Fig. ?.? shows the surface temperature and velocity field obtained from
185 MITgcm run at $\frac{1}{6}^{\circ }$ horizontal resolution on a $lat-lon$
186 grid in which the pole has been rotated by 90$^{\circ }$ on to the equator
187 (to avoid the converging of meridian in northern latitudes). 21 vertical
188 levels are used in the vertical with a `lopped cell' representation of
189 topography. The development and propagation of anomalously warm and cold
190 eddies can be clearly been seen in the Gulf Stream region. The transport of
191 warm water northward by the mean flow of the Gulf Stream is also clearly
192 visible.
193
194 %% CNHbegin
195 \input{part1/ocean_gyres_figure}
196 %% CNHend
197
198
199 \subsection{Global ocean circulation}
200
201 Fig.E2a shows the pattern of ocean currents at the surface of a 4$^{\circ }$
202 global ocean model run with 15 vertical levels. Lopped cells are used to
203 represent topography on a regular $lat-lon$ grid extending from 70$^{\circ
204 }N $ to 70$^{\circ }S$. The model is driven using monthly-mean winds with
205 mixed boundary conditions on temperature and salinity at the surface. The
206 transfer properties of ocean eddies, convection and mixing is parameterized
207 in this model.
208
209 Fig.E2b shows the meridional overturning circulation of the global ocean in
210 Sverdrups.
211
212 %%CNHbegin
213 \input{part1/global_circ_figure}
214 %%CNHend
215
216 \subsection{Convection and mixing over topography}
217
218 Dense plumes generated by localized cooling on the continental shelf of the
219 ocean may be influenced by rotation when the deformation radius is smaller
220 than the width of the cooling region. Rather than gravity plumes, the
221 mechanism for moving dense fluid down the shelf is then through geostrophic
222 eddies. The simulation shown in the figure (blue is cold dense fluid, red is
223 warmer, lighter fluid) employs the non-hydrostatic capability of MITgcm to
224 trigger convection by surface cooling. The cold, dense water falls down the
225 slope but is deflected along the slope by rotation. It is found that
226 entrainment in the vertical plane is reduced when rotational control is
227 strong, and replaced by lateral entrainment due to the baroclinic
228 instability of the along-slope current.
229
230 %%CNHbegin
231 \input{part1/convect_and_topo}
232 %%CNHend
233
234 \subsection{Boundary forced internal waves}
235
236 The unique ability of MITgcm to treat non-hydrostatic dynamics in the
237 presence of complex geometry makes it an ideal tool to study internal wave
238 dynamics and mixing in oceanic canyons and ridges driven by large amplitude
239 barotropic tidal currents imposed through open boundary conditions.
240
241 Fig. ?.? shows the influence of cross-slope topographic variations on
242 internal wave breaking - the cross-slope velocity is in color, the density
243 contoured. The internal waves are excited by application of open boundary
244 conditions on the left.\ They propagate to the sloping boundary (represented
245 using MITgcm's finite volume spatial discretization) where they break under
246 nonhydrostatic dynamics.
247
248 %%CNHbegin
249 \input{part1/boundary_forced_waves}
250 %%CNHend
251
252 \subsection{Parameter sensitivity using the adjoint of MITgcm}
253
254 Forward and tangent linear counterparts of MITgcm are supported using an
255 `automatic adjoint compiler'. These can be used in parameter sensitivity and
256 data assimilation studies.
257
258 As one example of application of the MITgcm adjoint, Fig.E4 maps the
259 gradient $\frac{\partial J}{\partial \mathcal{H}}$where $J$ is the magnitude
260 of the overturning streamfunction shown in fig?.? at 40$^{\circ }$N and $%
261 \mathcal{H}$ is the air-sea heat flux 100 years before. We see that $J$ is
262 sensitive to heat fluxes over the Labrador Sea, one of the important sources
263 of deep water for the thermohaline circulations. This calculation also
264 yields sensitivities to all other model parameters.
265
266 %%CNHbegin
267 \input{part1/adj_hf_ocean_figure}
268 %%CNHend
269
270 \subsection{Global state estimation of the ocean}
271
272 An important application of MITgcm is in state estimation of the global
273 ocean circulation. An appropriately defined `cost function', which measures
274 the departure of the model from observations (both remotely sensed and
275 insitu) over an interval of time, is minimized by adjusting `control
276 parameters' such as air-sea fluxes, the wind field, the initial conditions
277 etc. Figure ?.? shows an estimate of the time-mean surface elevation of the
278 ocean obtained by bringing the model in to consistency with altimetric and
279 in-situ observations over the period 1992-1997.
280
281 %% CNHbegin
282 \input{part1/globes_figure}
283 %% CNHend
284
285 \subsection{Ocean biogeochemical cycles}
286
287 MITgcm is being used to study global biogeochemical cycles in the ocean. For
288 example one can study the effects of interannual changes in meteorological
289 forcing and upper ocean circulation on the fluxes of carbon dioxide and
290 oxygen between the ocean and atmosphere. The figure shows the annual air-sea
291 flux of oxygen and its relation to density outcrops in the southern oceans
292 from a single year of a global, interannually varying simulation.
293
294 %%CNHbegin
295 \input{part1/biogeo_figure}
296 %%CNHend
297
298 \subsection{Simulations of laboratory experiments}
299
300 Figure ?.? shows MITgcm being used to simulate a laboratory experiment
301 enquiring in to the dynamics of the Antarctic Circumpolar Current (ACC). An
302 initially homogeneous tank of water ($1m$ in diameter) is driven from its
303 free surface by a rotating heated disk. The combined action of mechanical
304 and thermal forcing creates a lens of fluid which becomes baroclinically
305 unstable. The stratification and depth of penetration of the lens is
306 arrested by its instability in a process analogous to that whic sets the
307 stratification of the ACC.
308
309 %%CNHbegin
310 \input{part1/lab_figure}
311 %%CNHend
312
313 % $Header: /u/u0/gcmpack/mitgcmdoc/part1/manual.tex,v 1.2 2001/10/09 10:48:03 cnh Exp $
314 % $Name: $
315
316 \section{Continuous equations in `r' coordinates}
317
318 To render atmosphere and ocean models from one dynamical core we exploit
319 `isomorphisms' between equation sets that govern the evolution of the
320 respective fluids - see fig.4%
321 \marginpar{
322 Fig.4. Isomorphisms}. One system of hydrodynamical equations is written down
323 and encoded. The model variables have different interpretations depending on
324 whether the atmosphere or ocean is being studied. Thus, for example, the
325 vertical coordinate `$r$' is interpreted as pressure, $p$, if we are
326 modeling the atmosphere and height, $z$, if we are modeling the ocean.
327
328 %%CNHbegin
329 \input{part1/zandpcoord_figure.tex}
330 %%CNHend
331
332 The state of the fluid at any time is characterized by the distribution of
333 velocity $\vec{\mathbf{v}}$, active tracers $\theta $ and $S$, a
334 `geopotential' $\phi $ and density $\rho =\rho (\theta ,S,p)$ which may
335 depend on $\theta $, $S$, and $p$. The equations that govern the evolution
336 of these fields, obtained by applying the laws of classical mechanics and
337 thermodynamics to a Boussinesq, Navier-Stokes fluid are, written in terms of
338 a generic vertical coordinate, $r$, see fig.5%
339 \marginpar{
340 Fig.5 The vertical coordinate of model}:
341
342 %%CNHbegin
343 \input{part1/vertcoord_figure.tex}
344 %%CNHend
345
346 \begin{equation*}
347 \frac{D\vec{\mathbf{v}_{h}}}{Dt}+\left( 2\vec{\Omega}\times \vec{\mathbf{v}}%
348 \right) _{h}+\mathbf{\nabla }_{h}\phi =\mathcal{F}_{\vec{\mathbf{v}_{h}}}%
349 \text{ horizontal mtm}
350 \end{equation*}
351
352 \begin{equation*}
353 \frac{D\dot{r}}{Dt}+\widehat{k}\cdot \left( 2\vec{\Omega}\times \vec{\mathbf{%
354 v}}\right) +\frac{\partial \phi }{\partial r}+b=\mathcal{F}_{\dot{r}}\text{
355 vertical mtm}
356 \end{equation*}
357
358 \begin{equation}
359 \mathbf{\nabla }_{h}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \dot{r}}{%
360 \partial r}=0\text{ continuity} \label{eq:continuous}
361 \end{equation}
362
363 \begin{equation*}
364 b=b(\theta ,S,r)\text{ equation of state}
365 \end{equation*}
366
367 \begin{equation*}
368 \frac{D\theta }{Dt}=\mathcal{Q}_{\theta }\text{ potential temperature}
369 \end{equation*}
370
371 \begin{equation*}
372 \frac{DS}{Dt}=\mathcal{Q}_{S}\text{ humidity/salinity}
373 \end{equation*}
374
375 Here:
376
377 \begin{equation*}
378 r\text{ is the vertical coordinate}
379 \end{equation*}
380
381 \begin{equation*}
382 \frac{D}{Dt}=\frac{\partial }{\partial t}+\vec{\mathbf{v}}\cdot \nabla \text{
383 is the total derivative}
384 \end{equation*}
385
386 \begin{equation*}
387 \mathbf{\nabla }=\mathbf{\nabla }_{h}+\widehat{k}\frac{\partial }{\partial r}%
388 \text{ is the `grad' operator}
389 \end{equation*}
390 with $\mathbf{\nabla }_{h}$ operating in the horizontal and $\widehat{k}%
391 \frac{\partial }{\partial r}$ operating in the vertical, where $\widehat{k}$
392 is a unit vector in the vertical
393
394 \begin{equation*}
395 t\text{ is time}
396 \end{equation*}
397
398 \begin{equation*}
399 \vec{\mathbf{v}}=(u,v,\dot{r})=(\vec{\mathbf{v}}_{h},\dot{r})\text{ is the
400 velocity}
401 \end{equation*}
402
403 \begin{equation*}
404 \phi \text{ is the `pressure'/`geopotential'}
405 \end{equation*}
406
407 \begin{equation*}
408 \vec{\Omega}\text{ is the Earth's rotation}
409 \end{equation*}
410
411 \begin{equation*}
412 b\text{ is the `buoyancy'}
413 \end{equation*}
414
415 \begin{equation*}
416 \theta \text{ is potential temperature}
417 \end{equation*}
418
419 \begin{equation*}
420 S\text{ is specific humidity in the atmosphere; salinity in the ocean}
421 \end{equation*}
422
423 \begin{equation*}
424 \mathcal{F}_{\vec{\mathbf{v}}}\text{ are forcing and dissipation of }\vec{%
425 \mathbf{v}}
426 \end{equation*}
427
428 \begin{equation*}
429 \mathcal{Q}_{\theta }\mathcal{\ }\text{are forcing and dissipation of }\theta
430 \end{equation*}
431
432 \begin{equation*}
433 \mathcal{Q}_{S}\mathcal{\ }\text{are forcing and dissipation of }S
434 \end{equation*}
435
436 The $\mathcal{F}^{\prime }s$ and $\mathcal{Q}^{\prime }s$ are provided by
437 extensive `physics' packages for atmosphere and ocean described in Chapter 6.
438
439 \subsection{Kinematic Boundary conditions}
440
441 \subsubsection{vertical}
442
443 at fixed and moving $r$ surfaces we set (see fig.5):
444
445 \begin{equation}
446 \dot{r}=0atr=R_{fixed}(x,y)\text{ (ocean bottom, top of the atmosphere)}
447 \label{eq:fixedbc}
448 \end{equation}
449
450 \begin{equation}
451 \dot{r}=\frac{Dr}{Dt}atr=R_{moving}\text{ \
452 (oceansurface,bottomoftheatmosphere)} \label{eq:movingbc}
453 \end{equation}
454
455 Here
456
457 \begin{equation*}
458 R_{moving}=R_{o}+\eta
459 \end{equation*}
460 where $R_{o}(x,y)$ is the `$r-$value' (height or pressure, depending on
461 whether we are in the atmosphere or ocean) of the `moving surface' in the
462 resting fluid and $\eta $ is the departure from $R_{o}(x,y)$ in the presence
463 of motion.
464
465 \subsubsection{horizontal}
466
467 \begin{equation}
468 \vec{\mathbf{v}}\cdot \vec{\mathbf{n}}=0 \label{eq:noflow}
469 \end{equation}%
470 where $\vec{\mathbf{n}}$ is the normal to a solid boundary.
471
472 \subsection{Atmosphere}
473
474 In the atmosphere, see fig.5, we interpret:
475
476 \begin{equation}
477 r=p\text{ is the pressure} \label{eq:atmos-r}
478 \end{equation}
479
480 \begin{equation}
481 \dot{r}=\frac{Dp}{Dt}=\omega \text{ is the vertical velocity in }p\text{
482 coordinates} \label{eq:atmos-omega}
483 \end{equation}
484
485 \begin{equation}
486 \phi =g\,z\text{ is the geopotential height} \label{eq:atmos-phi}
487 \end{equation}
488
489 \begin{equation}
490 b=\frac{\partial \Pi }{\partial p}\theta \text{ is the buoyancy}
491 \label{eq:atmos-b}
492 \end{equation}
493
494 \begin{equation}
495 \theta =T(\frac{p_{c}}{p})^{\kappa }\text{ is potential temperature}
496 \label{eq:atmos-theta}
497 \end{equation}
498
499 \begin{equation}
500 S=q,\text{ is the specific humidity} \label{eq:atmos-s}
501 \end{equation}
502 where
503
504 \begin{equation*}
505 T\text{ is absolute temperature}
506 \end{equation*}%
507 \begin{equation*}
508 p\text{ is the pressure}
509 \end{equation*}%
510 \begin{eqnarray*}
511 &&z\text{ is the height of the pressure surface} \\
512 &&g\text{ is the acceleration due to gravity}
513 \end{eqnarray*}
514
515 In the above the ideal gas law, $p=\rho RT$, has been expressed in terms of
516 the Exner function $\Pi (p)$ given by (see Appendix Atmosphere)
517 \begin{equation}
518 \Pi (p)=c_{p}(\frac{p}{p_{c}})^{\kappa } \label{eq:exner}
519 \end{equation}%
520 where $p_{c}$ is a reference pressure and $\kappa =R/c_{p}$ with $R$ the gas
521 constant and $c_{p}$ the specific heat of air at constant pressure.
522
523 At the top of the atmosphere (which is `fixed' in our $r$ coordinate):
524
525 \begin{equation*}
526 R_{fixed}=p_{top}=0
527 \end{equation*}
528 In a resting atmosphere the elevation of the mountains at the bottom is
529 given by
530 \begin{equation*}
531 R_{moving}=R_{o}(x,y)=p_{o}(x,y)
532 \end{equation*}
533 i.e. the (hydrostatic) pressure at the top of the mountains in a resting
534 atmosphere.
535
536 The boundary conditions at top and bottom are given by:
537
538 \begin{eqnarray}
539 &&\omega =0~\text{at }r=R_{fixed} \text{ (top of the atmosphere)}
540 \label{eq:fixed-bc-atmos} \\
541 \omega &=&\frac{Dp_{s}}{Dt}\text{; at }r=R_{moving}\text{ (bottom of the
542 atmosphere)} \label{eq:moving-bc-atmos}
543 \end{eqnarray}
544
545 Then the (hydrostatic form of) eq(\ref{eq:continuous}) yields a consistent
546 set of atmospheric equations which, for convenience, are written out in $p$
547 coordinates in Appendix Atmosphere - see eqs(\ref{eq:atmos-prime}).
548
549 \subsection{Ocean}
550
551 In the ocean we interpret:
552 \begin{eqnarray}
553 r &=&z\text{ is the height} \label{eq:ocean-z} \\
554 \dot{r} &=&\frac{Dz}{Dt}=w\text{ is the vertical velocity}
555 \label{eq:ocean-w} \\
556 \phi &=&\frac{p}{\rho _{c}}\text{ is the pressure} \label{eq:ocean-p} \\
557 b(\theta ,S,r) &=&\frac{g}{\rho _{c}}\left( \rho (\theta ,S,r)-\rho
558 _{c}\right) \text{ is the buoyancy} \label{eq:ocean-b}
559 \end{eqnarray}
560 where $\rho _{c}$ is a fixed reference density of water and $g$ is the
561 acceleration due to gravity.\noindent
562
563 In the above
564
565 At the bottom of the ocean: $R_{fixed}(x,y)=-H(x,y)$.
566
567 The surface of the ocean is given by: $R_{moving}=\eta $
568
569 The position of the resting free surface of the ocean is given by $%
570 R_{o}=Z_{o}=0$.
571
572 Boundary conditions are:
573
574 \begin{eqnarray}
575 w &=&0~\text{at }r=R_{fixed}\text{ (ocean bottom)} \label{eq:fixed-bc-ocean}
576 \\
577 w &=&\frac{D\eta }{Dt}\text{ at }r=R_{moving}=\eta \text{ (ocean surface) %
578 \label{eq:moving-bc-ocean}}
579 \end{eqnarray}
580 where $\eta $ is the elevation of the free surface.
581
582 Then eq(\ref{eq:continuous}) yields a consistent set of oceanic equations
583 which, for convenience, are written out in $z$ coordinates in Appendix Ocean
584 - see eqs(\ref{eq:ocean-mom}) to (\ref{eq:ocean-salt}).
585
586 \subsection{Hydrostatic, Quasi-hydrostatic, Quasi-nonhydrostatic and
587 Non-hydrostatic forms}
588
589 Let us separate $\phi $ in to surface, hydrostatic and non-hydrostatic terms:
590
591 \begin{equation}
592 \phi (x,y,r)=\phi _{s}(x,y)+\phi _{hyd}(x,y,r)+\phi _{nh}(x,y,r)
593 \label{eq:phi-split}
594 \end{equation}%
595 and write eq(\ref{incompressible}a,b) in the form:
596
597 \begin{equation}
598 \frac{\partial \vec{\mathbf{v}_{h}}}{\partial t}+\mathbf{\nabla }_{h}\phi
599 _{s}+\mathbf{\nabla }_{h}\phi _{hyd}+\epsilon _{nh}\mathbf{\nabla }_{h}\phi
600 _{nh}=\vec{\mathbf{G}}_{\vec{v}_{h}} \label{eq:mom-h}
601 \end{equation}
602
603 \begin{equation}
604 \frac{\partial \phi _{hyd}}{\partial r}=-b \label{eq:hydrostatic}
605 \end{equation}
606
607 \begin{equation}
608 \epsilon _{nh}\frac{\partial \dot{r}}{\partial t}+\frac{\partial \phi _{nh}}{%
609 \partial r}=G_{\dot{r}} \label{eq:mom-w}
610 \end{equation}
611 Here $\epsilon _{nh}$ is a non-hydrostatic parameter.
612
613 The $\left( \vec{\mathbf{G}}_{\vec{v}},G_{\dot{r}}\right) $ in eq(\ref%
614 {eq:mom-h}) and (\ref{eq:mom-w}) represent advective, metric and Coriolis
615 terms in the momentum equations. In spherical coordinates they take the form%
616 \footnote{%
617 In the hydrostatic primitive equations (\textbf{HPE}) all underlined terms
618 in (\ref{eq:gu-speherical}), (\ref{eq:gv-spherical}) and (\ref%
619 {eq:gw-spherical}) are omitted; the singly-underlined terms are included in
620 the quasi-hydrostatic model (\textbf{QH}). The fully non-hydrostatic model (%
621 \textbf{NH}) includes all terms.} - see Marshall et al 1997a for a full
622 discussion:
623
624 \begin{equation}
625 \left.
626 \begin{tabular}{l}
627 $G_{u}=-\vec{\mathbf{v}}.\nabla u$ \\
628 $-\left\{ \underline{\frac{u\dot{r}}{{r}}}-\frac{uv\tan lat}{{r}}\right\} $
629 \\
630 $-\left\{ -2\Omega v\sin lat+\underline{2\Omega \dot{r}\cos lat}\right\} $
631 \\
632 $+\mathcal{F}_{u}$%
633 \end{tabular}%
634 \ \right\} \left\{
635 \begin{tabular}{l}
636 \textit{advection} \\
637 \textit{metric} \\
638 \textit{Coriolis} \\
639 \textit{\ Forcing/Dissipation}%
640 \end{tabular}%
641 \ \right. \qquad \label{eq:gu-speherical}
642 \end{equation}
643
644 \begin{equation}
645 \left.
646 \begin{tabular}{l}
647 $G_{v}=-\vec{\mathbf{v}}.\nabla v$ \\
648 $-\left\{ \underline{\frac{v\dot{r}}{{r}}}-\frac{u^{2}\tan lat}{{r}}\right\}
649 $ \\
650 $-\left\{ -2\Omega u\sin lat\right\} $ \\
651 $+\mathcal{F}_{v}$%
652 \end{tabular}%
653 \ \right\} \left\{
654 \begin{tabular}{l}
655 \textit{advection} \\
656 \textit{metric} \\
657 \textit{Coriolis} \\
658 \textit{\ Forcing/Dissipation}%
659 \end{tabular}%
660 \ \right. \qquad \label{eq:gv-spherical}
661 \end{equation}%
662 \qquad \qquad \qquad \qquad \qquad
663
664 \begin{equation}
665 \left.
666 \begin{tabular}{l}
667 $G_{\dot{r}}=-\underline{\underline{\vec{\mathbf{v}}.\nabla \dot{r}}}$ \\
668 $+\left\{ \underline{\frac{u^{_{^{2}}}+v^{2}}{{r}}}\right\} $ \\
669 ${+}\underline{{2\Omega u\cos lat}}$ \\
670 $\underline{\underline{\mathcal{F}_{\dot{r}}}}$%
671 \end{tabular}%
672 \ \right\} \left\{
673 \begin{tabular}{l}
674 \textit{advection} \\
675 \textit{metric} \\
676 \textit{Coriolis} \\
677 \textit{\ Forcing/Dissipation}%
678 \end{tabular}%
679 \ \right. \label{eq:gw-spherical}
680 \end{equation}%
681 \qquad \qquad \qquad \qquad \qquad
682
683 In the above `${r}$' is the distance from the center of the earth and `$lat$%
684 ' is latitude.
685
686 Grad and div operators in spherical coordinates are defined in appendix
687 OPERATORS.%
688 \marginpar{
689 Fig.6 Spherical polar coordinate system.}
690
691 %%CNHbegin
692 \input{part1/sphere_coord_figure.tex}
693 %%CNHend
694
695 \subsubsection{Shallow atmosphere approximation}
696
697 Most models are based on the `hydrostatic primitive equations' (HPE's) in
698 which the vertical momentum equation is reduced to a statement of
699 hydrostatic balance and the `traditional approximation' is made in which the
700 Coriolis force is treated approximately and the shallow atmosphere
701 approximation is made.\ The MITgcm need not make the `traditional
702 approximation'. To be able to support consistent non-hydrostatic forms the
703 shallow atmosphere approximation can be relaxed - when dividing through by $%
704 r $ in, for example, (\ref{eq:gu-speherical}), we do not replace $r$ by $a$,
705 the radius of the earth.
706
707 \subsubsection{Hydrostatic and quasi-hydrostatic forms}
708
709 These are discussed at length in Marshall et al (1997a).
710
711 In the `hydrostatic primitive equations' (\textbf{HPE)} all the underlined
712 terms in Eqs. (\ref{eq:gu-speherical} $\rightarrow $\ \ref{eq:gw-spherical})
713 are neglected and `${r}$' is replaced by `$a$', the mean radius of the
714 earth. Once the pressure is found at one level - e.g. by inverting a 2-d
715 Elliptic equation for $\phi _{s}$ at $r=R_{moving}$ - the pressure can be
716 computed at all other levels by integration of the hydrostatic relation, eq(%
717 \ref{eq:hydrostatic}).
718
719 In the `quasi-hydrostatic' equations (\textbf{QH)} strict balance between
720 gravity and vertical pressure gradients is not imposed. The $2\Omega u\cos
721 \phi $ Coriolis term are not neglected and are balanced by a non-hydrostatic
722 contribution to the pressure field: only the terms underlined twice in Eqs. (%
723 \ref{eq:gu-speherical}$\rightarrow $\ \ref{eq:gw-spherical}) are set to zero
724 and, simultaneously, the shallow atmosphere approximation is relaxed. In
725 \textbf{QH}\ \textit{all} the metric terms are retained and the full
726 variation of the radial position of a particle monitored. The \textbf{QH}\
727 vertical momentum equation (\ref{eq:mom-w}) becomes:
728
729 \begin{equation*}
730 \frac{\partial \phi _{nh}}{\partial r}=2\Omega u\cos lat
731 \end{equation*}
732 making a small correction to the hydrostatic pressure.
733
734 \textbf{QH} has good energetic credentials - they are the same as for
735 \textbf{HPE}. Importantly, however, it has the same angular momentum
736 principle as the full non-hydrostatic model (\textbf{NH)} - see Marshall
737 et.al., 1997a. As in \textbf{HPE }only a 2-d elliptic problem need be solved.
738
739 \subsubsection{Non-hydrostatic and quasi-nonhydrostatic forms}
740
741 The MIT model presently supports a full non-hydrostatic ocean isomorph, but
742 only a quasi-non-hydrostatic atmospheric isomorph.
743
744 \paragraph{Non-hydrostatic Ocean}
745
746 In the non-hydrostatic ocean model all terms in equations Eqs.(\ref%
747 {eq:gu-speherical} $\rightarrow $\ \ref{eq:gw-spherical}) are retained. A
748 three dimensional elliptic equation must be solved subject to Neumann
749 boundary conditions (see below). It is important to note that use of the
750 full \textbf{NH} does not admit any new `fast' waves in to the system - the
751 incompressible condition eq(\ref{eq:continuous})c has already filtered out
752 acoustic modes. It does, however, ensure that the gravity waves are treated
753 accurately with an exact dispersion relation. The \textbf{NH} set has a
754 complete angular momentum principle and consistent energetics - see White
755 and Bromley, 1995; Marshall et.al.\ 1997a.
756
757 \paragraph{Quasi-nonhydrostatic Atmosphere}
758
759 In the non-hydrostatic version of our atmospheric model we approximate $\dot{%
760 r}$ in the vertical momentum eqs(\ref{eq:mom-w}) and (\ref{eq:gv-spherical})
761 (but only here) by:
762
763 \begin{equation}
764 \dot{r}=\frac{Dp}{Dt}=\frac{1}{g}\frac{D\phi }{Dt} \label{eq:quasi-nh-w}
765 \end{equation}%
766 where $p_{hy}$ is the hydrostatic pressure.
767
768 \subsubsection{Summary of equation sets supported by model}
769
770 \paragraph{Atmosphere}
771
772 Hydrostatic, and quasi-hydrostatic and quasi non-hydrostatic forms of the
773 compressible non-Boussinesq equations in $p-$coordinates are supported.
774
775 \subparagraph{Hydrostatic and quasi-hydrostatic}
776
777 The hydrostatic set is written out in $p-$coordinates in appendix Atmosphere
778 - see eq(\ref{eq:atmos-prime}).
779
780 \subparagraph{Quasi-nonhydrostatic}
781
782 A quasi-nonhydrostatic form is also supported.
783
784 \paragraph{Ocean}
785
786 \subparagraph{Hydrostatic and quasi-hydrostatic}
787
788 Hydrostatic, and quasi-hydrostatic forms of the incompressible Boussinesq
789 equations in $z-$coordinates are supported.
790
791 \subparagraph{Non-hydrostatic}
792
793 Non-hydrostatic forms of the incompressible Boussinesq equations in $z-$%
794 coordinates are supported - see eqs(\ref{eq:ocean-mom}) to (\ref%
795 {eq:ocean-salt}).
796
797 \subsection{Solution strategy}
798
799 The method of solution employed in the \textbf{HPE}, \textbf{QH} and \textbf{%
800 NH} models is summarized in Fig.7.%
801 \marginpar{
802 Fig.7 Solution strategy} Under all dynamics, a 2-d elliptic equation is
803 first solved to find the surface pressure and the hydrostatic pressure at
804 any level computed from the weight of fluid above. Under \textbf{HPE} and
805 \textbf{QH} dynamics, the horizontal momentum equations are then stepped
806 forward and $\dot{r}$ found from continuity. Under \textbf{NH} dynamics a
807 3-d elliptic equation must be solved for the non-hydrostatic pressure before
808 stepping forward the horizontal momentum equations; $\dot{r}$ is found by
809 stepping forward the vertical momentum equation.
810
811 %%CNHbegin
812 \input{part1/solution_strategy_figure.tex}
813 %%CNHend
814
815 There is no penalty in implementing \textbf{QH} over \textbf{HPE} except, of
816 course, some complication that goes with the inclusion of $\cos \phi \ $%
817 Coriolis terms and the relaxation of the shallow atmosphere approximation.
818 But this leads to negligible increase in computation. In \textbf{NH}, in
819 contrast, one additional elliptic equation - a three-dimensional one - must
820 be inverted for $p_{nh}$. However the `overhead' of the \textbf{NH} model is
821 essentially negligible in the hydrostatic limit (see detailed discussion in
822 Marshall et al, 1997) resulting in a non-hydrostatic algorithm that, in the
823 hydrostatic limit, is as computationally economic as the \textbf{HPEs}.
824
825 \subsection{Finding the pressure field}
826
827 Unlike the prognostic variables $u$, $v$, $w$, $\theta $ and $S$, the
828 pressure field must be obtained diagnostically. We proceed, as before, by
829 dividing the total (pressure/geo) potential in to three parts, a surface
830 part, $\phi _{s}(x,y)$, a hydrostatic part $\phi _{hyd}(x,y,r)$ and a
831 non-hydrostatic part $\phi _{nh}(x,y,r)$, as in (\ref{eq:phi-split}), and
832 writing the momentum equation as in (\ref{eq:mom-h}).
833
834 \subsubsection{Hydrostatic pressure}
835
836 Hydrostatic pressure is obtained by integrating (\ref{eq:hydrostatic})
837 vertically from $r=R_{o}$ where $\phi _{hyd}(r=R_{o})=0$, to yield:
838
839 \begin{equation*}
840 \int_{r}^{R_{o}}\frac{\partial \phi _{hyd}}{\partial r}dr=\left[ \phi _{hyd}%
841 \right] _{r}^{R_{o}}=\int_{r}^{R_{o}}-bdr
842 \end{equation*}
843 and so
844
845 \begin{equation}
846 \phi _{hyd}(x,y,r)=\int_{r}^{R_{o}}bdr \label{eq:hydro-phi}
847 \end{equation}
848
849 The model can be easily modified to accommodate a loading term (e.g
850 atmospheric pressure pushing down on the ocean's surface) by setting:
851
852 \begin{equation}
853 \phi _{hyd}(r=R_{o})=loading \label{eq:loading}
854 \end{equation}
855
856 \subsubsection{Surface pressure}
857
858 The surface pressure equation can be obtained by integrating continuity, (%
859 \ref{eq:continuous})c, vertically from $r=R_{fixed}$ to $r=R_{moving}$
860
861 \begin{equation*}
862 \int_{R_{fixed}}^{R_{moving}}\left( \mathbf{\nabla }_{h}\cdot \vec{\mathbf{v}%
863 }_{h}+\partial _{r}\dot{r}\right) dr=0
864 \end{equation*}
865
866 Thus:
867
868 \begin{equation*}
869 \frac{\partial \eta }{\partial t}+\vec{\mathbf{v}}.\nabla \eta
870 +\int_{R_{fixed}}^{R_{moving}}\mathbf{\nabla }_{h}\cdot \vec{\mathbf{v}}%
871 _{h}dr=0
872 \end{equation*}
873 where $\eta =R_{moving}-R_{o}$ is the free-surface $r$-anomaly in units of $%
874 r $. The above can be rearranged to yield, using Leibnitz's theorem:
875
876 \begin{equation}
877 \frac{\partial \eta }{\partial t}+\mathbf{\nabla }_{h}\cdot
878 \int_{R_{fixed}}^{R_{moving}}\vec{\mathbf{v}}_{h}dr=\text{source}
879 \label{eq:free-surface}
880 \end{equation}%
881 where we have incorporated a source term.
882
883 Whether $\phi $ is pressure (ocean model, $p/\rho _{c}$) or geopotential
884 (atmospheric model), in (\ref{mtm-split}), the horizontal gradient term can
885 be written
886 \begin{equation}
887 \mathbf{\nabla }_{h}\phi _{s}=\mathbf{\nabla }_{h}\left( b_{s}\eta \right)
888 \label{eq:phi-surf}
889 \end{equation}%
890 where $b_{s}$ is the buoyancy at the surface.
891
892 In the hydrostatic limit ($\epsilon _{nh}=0$), Eqs(\ref{eq:mom-h}), (\ref%
893 {eq:free-surface}) and (\ref{eq:phi-surf}) can be solved by inverting a 2-d
894 elliptic equation for $\phi _{s}$ as described in Chapter 2. Both `free
895 surface' and `rigid lid' approaches are available.
896
897 \subsubsection{Non-hydrostatic pressure}
898
899 Taking the horizontal divergence of (\ref{hor-mtm}) and adding $\frac{%
900 \partial }{\partial r}$ of (\ref{vertmtm}), invoking the continuity equation
901 (\ref{incompressible}), we deduce that:
902
903 \begin{equation}
904 \nabla _{3}^{2}\phi _{nh}=\nabla .\vec{\mathbf{G}}_{\vec{v}}-\left( \mathbf{%
905 \nabla }_{h}^{2}\phi _{s}+\mathbf{\nabla }^{2}\phi _{hyd}\right) =\nabla .%
906 \vec{\mathbf{F}} \label{eq:3d-invert}
907 \end{equation}
908
909 For a given rhs this 3-d elliptic equation must be inverted for $\phi _{nh}$
910 subject to appropriate choice of boundary conditions. This method is usually
911 called \textit{The Pressure Method} [Harlow and Welch, 1965; Williams, 1969;
912 Potter, 1976]. In the hydrostatic primitive equations case (\textbf{HPE}),
913 the 3-d problem does not need to be solved.
914
915 \paragraph{Boundary Conditions}
916
917 We apply the condition of no normal flow through all solid boundaries - the
918 coasts (in the ocean) and the bottom:
919
920 \begin{equation}
921 \vec{\mathbf{v}}.\widehat{n}=0 \label{nonormalflow}
922 \end{equation}
923 where $\widehat{n}$ is a vector of unit length normal to the boundary. The
924 kinematic condition (\ref{nonormalflow}) is also applied to the vertical
925 velocity at $r=R_{moving}$. No-slip $\left( v_{T}=0\right) \ $or slip $%
926 \left( \partial v_{T}/\partial n=0\right) \ $conditions are employed on the
927 tangential component of velocity, $v_{T}$, at all solid boundaries,
928 depending on the form chosen for the dissipative terms in the momentum
929 equations - see below.
930
931 Eq.(\ref{nonormalflow}) implies, making use of (\ref{mtm-split}), that:
932
933 \begin{equation}
934 \widehat{n}.\nabla \phi _{nh}=\widehat{n}.\vec{\mathbf{F}}
935 \label{eq:inhom-neumann-nh}
936 \end{equation}
937 where
938
939 \begin{equation*}
940 \vec{\mathbf{F}}=\vec{\mathbf{G}}_{\vec{v}}-\left( \mathbf{\nabla }_{h}\phi
941 _{s}+\mathbf{\nabla }\phi _{hyd}\right)
942 \end{equation*}%
943 presenting inhomogeneous Neumann boundary conditions to the Elliptic problem
944 (\ref{eq:3d-invert}). As shown, for example, by Williams (1969), one can
945 exploit classical 3D potential theory and, by introducing an appropriately
946 chosen $\delta $-function sheet of `source-charge', replace the
947 inhomogeneous boundary condition on pressure by a homogeneous one. The
948 source term $rhs$ in (\ref{eq:3d-invert}) is the divergence of the vector $%
949 \vec{\mathbf{F}}.$ By simultaneously setting $%
950 \begin{array}{l}
951 \widehat{n}.\vec{\mathbf{F}}%
952 \end{array}%
953 =0$\ and $\widehat{n}.\nabla \phi _{nh}=0\ $on the boundary the following
954 self-consistent but simpler homogenized Elliptic problem is obtained:
955
956 \begin{equation*}
957 \nabla ^{2}\phi _{nh}=\nabla .\widetilde{\vec{\mathbf{F}}}\qquad
958 \end{equation*}%
959 where $\widetilde{\vec{\mathbf{F}}}$ is a modified $\vec{\mathbf{F}}$ such
960 that $\widetilde{\vec{\mathbf{F}}}.\widehat{n}=0$. As is implied by (\ref%
961 {eq:inhom-neumann-nh}) the modified boundary condition becomes:
962
963 \begin{equation}
964 \widehat{n}.\nabla \phi _{nh}=0 \label{eq:hom-neumann-nh}
965 \end{equation}
966
967 If the flow is `close' to hydrostatic balance then the 3-d inversion
968 converges rapidly because $\phi _{nh}\ $is then only a small correction to
969 the hydrostatic pressure field (see the discussion in Marshall et al, a,b).
970
971 The solution $\phi _{nh}\ $to (\ref{eq:3d-invert}) and (\ref{homneuman})
972 does not vanish at $r=R_{moving}$, and so refines the pressure there.
973
974 \subsection{Forcing/dissipation}
975
976 \subsubsection{Forcing}
977
978 The forcing terms $\mathcal{F}$ on the rhs of the equations are provided by
979 `physics packages' described in detail in chapter ??.
980
981 \subsubsection{Dissipation}
982
983 \paragraph{Momentum}
984
985 Many forms of momentum dissipation are available in the model. Laplacian and
986 biharmonic frictions are commonly used:
987
988 \begin{equation}
989 D_{V}=A_{h}\nabla _{h}^{2}v+A_{v}\frac{\partial ^{2}v}{\partial z^{2}}%
990 +A_{4}\nabla _{h}^{4}v \label{eq:dissipation}
991 \end{equation}
992 where $A_{h}$ and $A_{v}\ $are (constant) horizontal and vertical viscosity
993 coefficients and $A_{4}\ $is the horizontal coefficient for biharmonic
994 friction. These coefficients are the same for all velocity components.
995
996 \paragraph{Tracers}
997
998 The mixing terms for the temperature and salinity equations have a similar
999 form to that of momentum except that the diffusion tensor can be
1000 non-diagonal and have varying coefficients. $\qquad $%
1001 \begin{equation}
1002 D_{T,S}=\nabla .[\underline{\underline{K}}\nabla (T,S)]+K_{4}\nabla
1003 _{h}^{4}(T,S) \label{eq:diffusion}
1004 \end{equation}
1005 where $\underline{\underline{K}}\ $is the diffusion tensor and the $K_{4}\ $%
1006 horizontal coefficient for biharmonic diffusion. In the simplest case where
1007 the subgrid-scale fluxes of heat and salt are parameterized with constant
1008 horizontal and vertical diffusion coefficients, $\underline{\underline{K}}$,
1009 reduces to a diagonal matrix with constant coefficients:
1010
1011 \begin{equation}
1012 \qquad \qquad \qquad \qquad K=\left(
1013 \begin{array}{ccc}
1014 K_{h} & 0 & 0 \\
1015 0 & K_{h} & 0 \\
1016 0 & 0 & K_{v}%
1017 \end{array}
1018 \right) \qquad \qquad \qquad \label{eq:diagonal-diffusion-tensor}
1019 \end{equation}
1020 where $K_{h}\ $and $K_{v}\ $are the horizontal and vertical diffusion
1021 coefficients. These coefficients are the same for all tracers (temperature,
1022 salinity ... ).
1023
1024 \subsection{Vector invariant form}
1025
1026 For some purposes it is advantageous to write momentum advection in eq(\ref%
1027 {hor-mtm}) and (\ref{vertmtm}) in the (so-called) `vector invariant' form:
1028
1029 \begin{equation}
1030 \frac{D\vec{\mathbf{v}}}{Dt}=\frac{\partial \vec{\mathbf{v}}}{\partial t}%
1031 +\left( \nabla \times \vec{\mathbf{v}}\right) \times \vec{\mathbf{v}}+\nabla %
1032 \left[ \frac{1}{2}(\vec{\mathbf{v}}\cdot \vec{\mathbf{v}})\right]
1033 \label{eq:vi-identity}
1034 \end{equation}%
1035 This permits alternative numerical treatments of the non-linear terms based
1036 on their representation as a vorticity flux. Because gradients of coordinate
1037 vectors no longer appear on the rhs of (\ref{eq:vi-identity}), explicit
1038 representation of the metric terms in (\ref{eq:gu-speherical}), (\ref%
1039 {eq:gv-spherical}) and (\ref{eq:gw-spherical}), can be avoided: information
1040 about the geometry is contained in the areas and lengths of the volumes used
1041 to discretize the model.
1042
1043 \subsection{Adjoint}
1044
1045 Tangent linear and adjoint counterparts of the forward model and described
1046 in Chapter 5.
1047
1048 % $Header: /u/u0/gcmpack/mitgcmdoc/part1/manual.tex,v 1.2 2001/10/09 10:48:03 cnh Exp $
1049 % $Name: $
1050
1051 \section{Appendix ATMOSPHERE}
1052
1053 \subsection{Hydrostatic Primitive Equations for the Atmosphere in pressure
1054 coordinates}
1055
1056 \label{sect-hpe-p}
1057
1058 The hydrostatic primitive equations (HPEs) in p-coordinates are:
1059 \begin{eqnarray}
1060 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}%
1061 _{h}+\mathbf{\nabla }_{p}\phi &=&\vec{\mathbf{\mathcal{F}}}
1062 \label{eq:atmos-mom} \\
1063 \frac{\partial \phi }{\partial p}+\alpha &=&0 \label{eq-p-hydro-start} \\
1064 \mathbf{\nabla }_{p}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \omega }{%
1065 \partial p} &=&0 \label{eq:atmos-cont} \\
1066 p\alpha &=&RT \label{eq:atmos-eos} \\
1067 c_{v}\frac{DT}{Dt}+p\frac{D\alpha }{Dt} &=&\mathcal{Q} \label{eq:atmos-heat}
1068 \end{eqnarray}%
1069 where $\vec{\mathbf{v}}_{h}=(u,v,0)$ is the `horizontal' (on pressure
1070 surfaces) component of velocity,$\frac{D}{Dt}=\vec{\mathbf{v}}_{h}\cdot
1071 \mathbf{\nabla }_{p}+\omega \frac{\partial }{\partial p}$ is the total
1072 derivative, $f=2\Omega \sin lat$ is the Coriolis parameter, $\phi =gz$ is
1073 the geopotential, $\alpha =1/\rho $ is the specific volume, $\omega =\frac{Dp%
1074 }{Dt}$ is the vertical velocity in the $p-$coordinate. Equation(\ref%
1075 {eq:atmos-heat}) is the first law of thermodynamics where internal energy $%
1076 e=c_{v}T$, $T$ is temperature, $Q$ is the rate of heating per unit mass and $%
1077 p\frac{D\alpha }{Dt}$ is the work done by the fluid in compressing.
1078
1079 It is convenient to cast the heat equation in terms of potential temperature
1080 $\theta $ so that it looks more like a generic conservation law.
1081 Differentiating (\ref{eq:atmos-eos}) we get:
1082 \begin{equation*}
1083 p\frac{D\alpha }{Dt}+\alpha \frac{Dp}{Dt}=R\frac{DT}{Dt}
1084 \end{equation*}%
1085 which, when added to the heat equation (\ref{eq:atmos-heat}) and using $%
1086 c_{p}=c_{v}+R$, gives:
1087 \begin{equation}
1088 c_{p}\frac{DT}{Dt}-\alpha \frac{Dp}{Dt}=\mathcal{Q}
1089 \label{eq-p-heat-interim}
1090 \end{equation}%
1091 Potential temperature is defined:
1092 \begin{equation}
1093 \theta =T(\frac{p_{c}}{p})^{\kappa } \label{eq:potential-temp}
1094 \end{equation}%
1095 where $p_{c}$ is a reference pressure and $\kappa =R/c_{p}$. For convenience
1096 we will make use of the Exner function $\Pi (p)$ which defined by:
1097 \begin{equation}
1098 \Pi (p)=c_{p}(\frac{p}{p_{c}})^{\kappa } \label{Exner}
1099 \end{equation}%
1100 The following relations will be useful and are easily expressed in terms of
1101 the Exner function:
1102 \begin{equation*}
1103 c_{p}T=\Pi \theta \;\;;\;\;\frac{\partial \Pi }{\partial p}=\frac{\kappa \Pi
1104 }{p}\;\;;\;\;\alpha =\frac{\kappa \Pi \theta }{p}=\frac{\partial \ \Pi }{%
1105 \partial p}\theta \;\;;\;\;\frac{D\Pi }{Dt}=\frac{\partial \Pi }{\partial p}%
1106 \frac{Dp}{Dt}
1107 \end{equation*}%
1108 where $b=\frac{\partial \ \Pi }{\partial p}\theta $ is the buoyancy.
1109
1110 The heat equation is obtained by noting that
1111 \begin{equation*}
1112 c_{p}\frac{DT}{Dt}=\frac{D(\Pi \theta )}{Dt}=\Pi \frac{D\theta }{Dt}+\theta
1113 \frac{D\Pi }{Dt}=\Pi \frac{D\theta }{Dt}+\alpha \frac{Dp}{Dt}
1114 \end{equation*}
1115 and on substituting into (\ref{eq-p-heat-interim}) gives:
1116 \begin{equation}
1117 \Pi \frac{D\theta }{Dt}=\mathcal{Q}
1118 \label{eq:potential-temperature-equation}
1119 \end{equation}
1120 which is in conservative form.
1121
1122 For convenience in the model we prefer to step forward (\ref%
1123 {eq:potential-temperature-equation}) rather than (\ref{eq:atmos-heat}).
1124
1125 \subsubsection{Boundary conditions}
1126
1127 The upper and lower boundary conditions are :
1128 \begin{eqnarray}
1129 \mbox{at the top:}\;\;p=0 &&\text{, }\omega =\frac{Dp}{Dt}=0 \\
1130 \mbox{at the surface:}\;\;p=p_{s} &&\text{, }\phi =\phi _{topo}=g~Z_{topo}
1131 \label{eq:boundary-condition-atmosphere}
1132 \end{eqnarray}
1133 In $p$-coordinates, the upper boundary acts like a solid boundary ($\omega
1134 =0 $); in $z$-coordinates and the lower boundary is analogous to a free
1135 surface ($\phi $ is imposed and $\omega \neq 0$).
1136
1137 \subsubsection{Splitting the geo-potential}
1138
1139 For the purposes of initialization and reducing round-off errors, the model
1140 deals with perturbations from reference (or ``standard'') profiles. For
1141 example, the hydrostatic geopotential associated with the resting atmosphere
1142 is not dynamically relevant and can therefore be subtracted from the
1143 equations. The equations written in terms of perturbations are obtained by
1144 substituting the following definitions into the previous model equations:
1145 \begin{eqnarray}
1146 \theta &=&\theta _{o}+\theta ^{\prime } \label{eq:atmos-ref-prof-theta} \\
1147 \alpha &=&\alpha _{o}+\alpha ^{\prime } \label{eq:atmos-ref-prof-alpha} \\
1148 \phi &=&\phi _{o}+\phi ^{\prime } \label{eq:atmos-ref-prof-phi}
1149 \end{eqnarray}
1150 The reference state (indicated by subscript ``0'') corresponds to
1151 horizontally homogeneous atmosphere at rest ($\theta _{o},\alpha _{o},\phi
1152 _{o}$) with surface pressure $p_{o}(x,y)$ that satisfies $\phi
1153 _{o}(p_{o})=g~Z_{topo}$, defined:
1154 \begin{eqnarray*}
1155 \theta _{o}(p) &=&f^{n}(p) \\
1156 \alpha _{o}(p) &=&\Pi _{p}\theta _{o} \\
1157 \phi _{o}(p) &=&\phi _{topo}-\int_{p_{0}}^{p}\alpha _{o}dp
1158 \end{eqnarray*}
1159 %\begin{eqnarray*}
1160 %\phi'_\alpha & = & \int^p_{p_o} (\alpha_o -\alpha) dp \\
1161 %\phi'_s(x,y,t) & = & \int_{p_o}^{p_s} \alpha dp
1162 %\end{eqnarray*}
1163
1164 The final form of the HPE's in p coordinates is then:
1165 \begin{eqnarray}
1166 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}%
1167 _{h}+\mathbf{\nabla }_{p}\phi ^{\prime } &=&\vec{\mathbf{\mathcal{F}}} \\
1168 \frac{\partial \phi ^{\prime }}{\partial p}+\alpha ^{\prime } &=&0 \\
1169 \mathbf{\nabla }_{p}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \omega }{%
1170 \partial p} &=&0 \\
1171 \frac{\partial \Pi }{\partial p}\theta ^{\prime } &=&\alpha ^{\prime } \\
1172 \frac{D\theta }{Dt} &=&\frac{\mathcal{Q}}{\Pi } \label{eq:atmos-prime}
1173 \end{eqnarray}
1174
1175 % $Header: /u/u0/gcmpack/mitgcmdoc/part1/manual.tex,v 1.2 2001/10/09 10:48:03 cnh Exp $
1176 % $Name: $
1177
1178 \section{Appendix OCEAN}
1179
1180 \subsection{Equations of motion for the ocean}
1181
1182 We review here the method by which the standard (Boussinesq, incompressible)
1183 HPE's for the ocean written in z-coordinates are obtained. The
1184 non-Boussinesq equations for oceanic motion are:
1185 \begin{eqnarray}
1186 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}%
1187 _{h}+\frac{1}{\rho }\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}} \\
1188 \epsilon _{nh}\frac{Dw}{Dt}+g+\frac{1}{\rho }\frac{\partial p}{\partial z}
1189 &=&\epsilon _{nh}\mathcal{F}_{w} \\
1190 \frac{1}{\rho }\frac{D\rho }{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}%
1191 _{h}+\frac{\partial w}{\partial z} &=&0 \\
1192 \rho &=&\rho (\theta ,S,p) \\
1193 \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \\
1194 \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq:non-boussinesq}
1195 \end{eqnarray}%
1196 These equations permit acoustics modes, inertia-gravity waves,
1197 non-hydrostatic motions, a geostrophic (Rossby) mode and a thermo-haline
1198 mode. As written, they cannot be integrated forward consistently - if we
1199 step $\rho $ forward in (\ref{eq-zns-cont}), the answer will not be
1200 consistent with that obtained by stepping (\ref{eq-zns-heat}) and (\ref%
1201 {eq-zns-salt}) and then using (\ref{eq-zns-eos}) to yield $\rho $. It is
1202 therefore necessary to manipulate the system as follows. Differentiating the
1203 EOS (equation of state) gives:
1204
1205 \begin{equation}
1206 \frac{D\rho }{Dt}=\left. \frac{\partial \rho }{\partial \theta }\right|
1207 _{S,p}\frac{D\theta }{Dt}+\left. \frac{\partial \rho }{\partial S}\right|
1208 _{\theta ,p}\frac{DS}{Dt}+\left. \frac{\partial \rho }{\partial p}\right|
1209 _{\theta ,S}\frac{Dp}{Dt} \label{EOSexpansion}
1210 \end{equation}
1211
1212 Note that $\frac{\partial \rho }{\partial p}=\frac{1}{c_{s}^{2}}$ is the
1213 reciprocal of the sound speed ($c_{s}$) squared. Substituting into \ref%
1214 {eq-zns-cont} gives:
1215 \begin{equation}
1216 \frac{1}{\rho c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{%
1217 v}}+\partial _{z}w\approx 0 \label{eq-zns-pressure}
1218 \end{equation}
1219 where we have used an approximation sign to indicate that we have assumed
1220 adiabatic motion, dropping the $\frac{D\theta }{Dt}$ and $\frac{DS}{Dt}$.
1221 Replacing \ref{eq-zns-cont} with \ref{eq-zns-pressure} yields a system that
1222 can be explicitly integrated forward:
1223 \begin{eqnarray}
1224 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}%
1225 _{h}+\frac{1}{\rho }\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1226 \label{eq-cns-hmom} \\
1227 \epsilon _{nh}\frac{Dw}{Dt}+g+\frac{1}{\rho }\frac{\partial p}{\partial z}
1228 &=&\epsilon _{nh}\mathcal{F}_{w} \label{eq-cns-hydro} \\
1229 \frac{1}{\rho c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{%
1230 v}}_{h}+\frac{\partial w}{\partial z} &=&0 \label{eq-cns-cont} \\
1231 \rho &=&\rho (\theta ,S,p) \label{eq-cns-eos} \\
1232 \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-cns-heat} \\
1233 \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-cns-salt}
1234 \end{eqnarray}
1235
1236 \subsubsection{Compressible z-coordinate equations}
1237
1238 Here we linearize the acoustic modes by replacing $\rho $ with $\rho _{o}(z)$
1239 wherever it appears in a product (ie. non-linear term) - this is the
1240 `Boussinesq assumption'. The only term that then retains the full variation
1241 in $\rho $ is the gravitational acceleration:
1242 \begin{eqnarray}
1243 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}%
1244 _{h}+\frac{1}{\rho _{o}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1245 \label{eq-zcb-hmom} \\
1246 \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{o}}+\frac{1}{\rho _{o}}%
1247 \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1248 \label{eq-zcb-hydro} \\
1249 \frac{1}{\rho _{o}c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{%
1250 \mathbf{v}}_{h}+\frac{\partial w}{\partial z} &=&0 \label{eq-zcb-cont} \\
1251 \rho &=&\rho (\theta ,S,p) \label{eq-zcb-eos} \\
1252 \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-zcb-heat} \\
1253 \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-zcb-salt}
1254 \end{eqnarray}
1255 These equations still retain acoustic modes. But, because the
1256 ``compressible'' terms are linearized, the pressure equation \ref%
1257 {eq-zcb-cont} can be integrated implicitly with ease (the time-dependent
1258 term appears as a Helmholtz term in the non-hydrostatic pressure equation).
1259 These are the \emph{truly} compressible Boussinesq equations. Note that the
1260 EOS must have the same pressure dependency as the linearized pressure term,
1261 ie. $\left. \frac{\partial \rho }{\partial p}\right| _{\theta ,S}=\frac{1}{%
1262 c_{s}^{2}}$, for consistency.
1263
1264 \subsubsection{`Anelastic' z-coordinate equations}
1265
1266 The anelastic approximation filters the acoustic mode by removing the
1267 time-dependency in the continuity (now pressure-) equation (\ref{eq-zcb-cont}%
1268 ). This could be done simply by noting that $\frac{Dp}{Dt}\approx -g\rho _{o}%
1269 \frac{Dz}{Dt}=-g\rho _{o}w$, but this leads to an inconsistency between
1270 continuity and EOS. A better solution is to change the dependency on
1271 pressure in the EOS by splitting the pressure into a reference function of
1272 height and a perturbation:
1273 \begin{equation*}
1274 \rho =\rho (\theta ,S,p_{o}(z)+\epsilon _{s}p^{\prime })
1275 \end{equation*}
1276 Remembering that the term $\frac{Dp}{Dt}$ in continuity comes from
1277 differentiating the EOS, the continuity equation then becomes:
1278 \begin{equation*}
1279 \frac{1}{\rho _{o}c_{s}^{2}}\left( \frac{Dp_{o}}{Dt}+\epsilon _{s}\frac{%
1280 Dp^{\prime }}{Dt}\right) +\mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+%
1281 \frac{\partial w}{\partial z}=0
1282 \end{equation*}
1283 If the time- and space-scales of the motions of interest are longer than
1284 those of acoustic modes, then $\frac{Dp^{\prime }}{Dt}<<(\frac{Dp_{o}}{Dt},%
1285 \mathbf{\nabla }\cdot \vec{\mathbf{v}}_{h})$ in the continuity equations and
1286 $\left. \frac{\partial \rho }{\partial p}\right| _{\theta ,S}\frac{%
1287 Dp^{\prime }}{Dt}<<\left. \frac{\partial \rho }{\partial p}\right| _{\theta
1288 ,S}\frac{Dp_{o}}{Dt}$ in the EOS (\ref{EOSexpansion}). Thus we set $\epsilon
1289 _{s}=0$, removing the dependency on $p^{\prime }$ in the continuity equation
1290 and EOS. Expanding $\frac{Dp_{o}(z)}{Dt}=-g\rho _{o}w$ then leads to the
1291 anelastic continuity equation:
1292 \begin{equation}
1293 \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z}-%
1294 \frac{g}{c_{s}^{2}}w=0 \label{eq-za-cont1}
1295 \end{equation}
1296 A slightly different route leads to the quasi-Boussinesq continuity equation
1297 where we use the scaling $\frac{\partial \rho ^{\prime }}{\partial t}+%
1298 \mathbf{\nabla }_{3}\cdot \rho ^{\prime }\vec{\mathbf{v}}<<\mathbf{\nabla }%
1299 _{3}\cdot \rho _{o}\vec{\mathbf{v}}$ yielding:
1300 \begin{equation}
1301 \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{1}{\rho _{o}}\frac{%
1302 \partial \left( \rho _{o}w\right) }{\partial z}=0 \label{eq-za-cont2}
1303 \end{equation}
1304 Equations \ref{eq-za-cont1} and \ref{eq-za-cont2} are in fact the same
1305 equation if:
1306 \begin{equation}
1307 \frac{1}{\rho _{o}}\frac{\partial \rho _{o}}{\partial z}=\frac{-g}{c_{s}^{2}}
1308 \end{equation}
1309 Again, note that if $\rho _{o}$ is evaluated from prescribed $\theta _{o}$
1310 and $S_{o}$ profiles, then the EOS dependency on $p_{o}$ and the term $\frac{%
1311 g}{c_{s}^{2}}$ in continuity should be referred to those same profiles. The
1312 full set of `quasi-Boussinesq' or `anelastic' equations for the ocean are
1313 then:
1314 \begin{eqnarray}
1315 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}%
1316 _{h}+\frac{1}{\rho _{o}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1317 \label{eq-zab-hmom} \\
1318 \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{o}}+\frac{1}{\rho _{o}}%
1319 \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1320 \label{eq-zab-hydro} \\
1321 \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{1}{\rho _{o}}\frac{%
1322 \partial \left( \rho _{o}w\right) }{\partial z} &=&0 \label{eq-zab-cont} \\
1323 \rho &=&\rho (\theta ,S,p_{o}(z)) \label{eq-zab-eos} \\
1324 \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-zab-heat} \\
1325 \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-zab-salt}
1326 \end{eqnarray}
1327
1328 \subsubsection{Incompressible z-coordinate equations}
1329
1330 Here, the objective is to drop the depth dependence of $\rho _{o}$ and so,
1331 technically, to also remove the dependence of $\rho $ on $p_{o}$. This would
1332 yield the ``truly'' incompressible Boussinesq equations:
1333 \begin{eqnarray}
1334 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}%
1335 _{h}+\frac{1}{\rho _{c}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1336 \label{eq-ztb-hmom} \\
1337 \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{c}}+\frac{1}{\rho _{c}}%
1338 \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1339 \label{eq-ztb-hydro} \\
1340 \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z}
1341 &=&0 \label{eq-ztb-cont} \\
1342 \rho &=&\rho (\theta ,S) \label{eq-ztb-eos} \\
1343 \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-ztb-heat} \\
1344 \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-ztb-salt}
1345 \end{eqnarray}
1346 where $\rho _{c}$ is a constant reference density of water.
1347
1348 \subsubsection{Compressible non-divergent equations}
1349
1350 The above ``incompressible'' equations are incompressible in both the flow
1351 and the density. In many oceanic applications, however, it is important to
1352 retain compressibility effects in the density. To do this we must split the
1353 density thus:
1354 \begin{equation*}
1355 \rho =\rho _{o}+\rho ^{\prime }
1356 \end{equation*}%
1357 We then assert that variations with depth of $\rho _{o}$ are unimportant
1358 while the compressible effects in $\rho ^{\prime }$ are:
1359 \begin{equation*}
1360 \rho _{o}=\rho _{c}
1361 \end{equation*}%
1362 \begin{equation*}
1363 \rho ^{\prime }=\rho (\theta ,S,p_{o}(z))-\rho _{o}
1364 \end{equation*}%
1365 This then yields what we can call the semi-compressible Boussinesq
1366 equations:
1367 \begin{eqnarray}
1368 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}%
1369 _{h}+\frac{1}{\rho _{c}}\mathbf{\nabla }_{z}p^{\prime } &=&\vec{\mathbf{%
1370 \mathcal{F}}} \label{eq:ocean-mom} \\
1371 \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho ^{\prime }}{\rho _{c}}+\frac{1}{\rho
1372 _{c}}\frac{\partial p^{\prime }}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1373 \label{eq:ocean-wmom} \\
1374 \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z}
1375 &=&0 \label{eq:ocean-cont} \\
1376 \rho ^{\prime } &=&\rho (\theta ,S,p_{o}(z))-\rho _{c} \label{eq:ocean-eos}
1377 \\
1378 \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq:ocean-theta} \\
1379 \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq:ocean-salt}
1380 \end{eqnarray}%
1381 Note that the hydrostatic pressure of the resting fluid, including that
1382 associated with $\rho _{c}$, is subtracted out since it has no effect on the
1383 dynamics.
1384
1385 Though necessary, the assumptions that go into these equations are messy
1386 since we essentially assume a different EOS for the reference density and
1387 the perturbation density. Nevertheless, it is the hydrostatic ($\epsilon
1388 _{nh}=0$ form of these equations that are used throughout the ocean modeling
1389 community and referred to as the primitive equations (HPE).
1390
1391 % $Header: /u/u0/gcmpack/mitgcmdoc/part1/manual.tex,v 1.2 2001/10/09 10:48:03 cnh Exp $
1392 % $Name: $
1393
1394 \section{Appendix:OPERATORS}
1395
1396 \subsection{Coordinate systems}
1397
1398 \subsubsection{Spherical coordinates}
1399
1400 In spherical coordinates, the velocity components in the zonal, meridional
1401 and vertical direction respectively, are given by (see Fig.2) :
1402
1403 \begin{equation*}
1404 u=r\cos \phi \frac{D\lambda }{Dt}
1405 \end{equation*}
1406
1407 \begin{equation*}
1408 v=r\frac{D\phi }{Dt}\qquad
1409 \end{equation*}
1410 $\qquad \qquad \qquad \qquad $
1411
1412 \begin{equation*}
1413 \dot{r}=\frac{Dr}{Dt}
1414 \end{equation*}
1415
1416 Here $\phi $ is the latitude, $\lambda $ the longitude, $r$ the radial
1417 distance of the particle from the center of the earth, $\Omega $ is the
1418 angular speed of rotation of the Earth and $D/Dt$ is the total derivative.
1419
1420 The `grad' ($\nabla $) and `div' ($\nabla $.) operators are defined by, in
1421 spherical coordinates:
1422
1423 \begin{equation*}
1424 \nabla \equiv \left( \frac{1}{r\cos \phi }\frac{\partial }{\partial \lambda }%
1425 ,\frac{1}{r}\frac{\partial }{\partial \phi },\frac{\partial }{\partial r}%
1426 \right)
1427 \end{equation*}
1428
1429 \begin{equation*}
1430 \nabla .v\equiv \frac{1}{r\cos \phi }\left\{ \frac{\partial u}{\partial
1431 \lambda }+\frac{\partial }{\partial \phi }\left( v\cos \phi \right) \right\}
1432 +\frac{1}{r^{2}}\frac{\partial \left( r^{2}\dot{r}\right) }{\partial r}
1433 \end{equation*}
1434
1435 %%%% \end{document}

  ViewVC Help
Powered by ViewVC 1.1.22