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

Annotation of /manual/s_overview/text/manual.tex

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


Revision 1.2 - (hide annotations) (download) (as text)
Tue Oct 9 10:48:03 2001 UTC (23 years, 9 months ago) by cnh
Branch: MAIN
Changes since 1.1: +202 -239 lines
File MIME type: application/x-tex
Part1 updates

1 cnh 1.2 % $Header: /u/gcmpack/mitgcmdoc/manual.tex,v 1.3 2001/08/09 19:30:00 adcroft 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 cnh 1.1
22    
23     %%%% \documentclass[12pt]{book}
24 cnh 1.2 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
25 cnh 1.1 %%%% \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 cnh 1.2
34 cnh 1.1 %%%% %TCIDATA{OutputFilter=Latex.dll}
35 cnh 1.2 %%%% %TCIDATA{LastRevised=Thursday, October 04, 2001 14:41:22}
36 cnh 1.1 %%%% %TCIDATA{<META NAME="GraphicsSave" CONTENT="32">}
37     %%%% %TCIDATA{Language=American English}
38 cnh 1.2
39 cnh 1.1 %%%% \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 cnh 1.2 %%%%
54     %%%% \pagebreak
55 cnh 1.1
56 cnh 1.2 %%%% \part{MIT GCM basics}
57 cnh 1.1
58     % Section: Overview
59    
60     % $Header: /u/gcmpack/mitgcmdoc/part1/introduction.tex,v 1.1.1.1 2001/08/08 16:16:16 adcroft 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.1%
81     \marginpar{
82     Fig.1 One model}\ref{fig:onemodel}
83    
84     \item it has a non-hydrostatic capability and so can be used to study both
85     small-scale and large scale processes - see fig.2%
86     \marginpar{
87     Fig.2 All scales}\ref{fig:all-scales}
88    
89     \item finite volume techniques are employed yielding an intuitive
90     discretization and support for the treatment of irregular geometries using
91     orthogonal curvilinear grids and shaved cells - see fig.3%
92     \marginpar{
93     Fig.3 Finite volumes}\ref{fig:Finite volumes}
94    
95     \item tangent linear and adjoint counterparts are automatically maintained
96     along with the forward model, permitting sensitivity and optimization
97     studies.
98    
99     \item the model is developed to perform efficiently on a wide variety of
100     computational platforms.
101     \end{itemize}
102    
103     Key publications reporting on and charting the development of the model are
104     listed in an Appendix.
105    
106     We begin by briefly showing some of the results of the model in action to
107     give a feel for the wide range of problems that can be addressed using it.
108     \pagebreak
109    
110     % $Header: /u/gcmpack/mitgcmdoc/part1/illustration.tex,v 1.1.1.1 2001/08/08 16:16:16 adcroft Exp $
111     % $Name: $
112    
113     \section{Illustrations of the model in action}
114    
115     The MITgcm has been designed and used to model a wide range of phenomena,
116     from convection on the scale of meters in the ocean to the global pattern of
117     atmospheric winds - see fig.2\ref{fig:all-scales}. To give a flavor of the
118     kinds of problems the model has been used to study, we briefly describe some
119     of them here. A more detailed description of the underlying formulation,
120     numerical algorithm and implementation that lie behind these calculations is
121 cnh 1.2 given later. Indeed many of the illustrative examples shown below can be
122     easily reproduced: simply download the model (the minimum you need is a PC
123     running linux, together with a FORTRAN\ 77 compiler) and follow the examples
124     described in detail in the documentation.
125 cnh 1.1
126     \subsection{Global atmosphere: `Held-Suarez' benchmark}
127    
128 cnh 1.2 A novel feature of MITgcm is its ability to simulate both atmospheric and
129     oceanographic flows at both small and large scales.
130    
131     Fig.E1a.\ref{fig:Held-Suarez} shows an instantaneous plot of the 500$mb$
132     temperature field obtained using the atmospheric isomorph of MITgcm run at
133     2.8$^{\circ }$ resolution on the cubed sphere. We see cold air over the pole
134     (blue) and warm air along an equatorial band (red). Fully developed
135     baroclinic eddies spawned in the northern hemisphere storm track are
136     evident. There are no mountains or land-sea contrast in this calculation,
137     but you can easily put them in. The model is driven by relaxation to a
138     radiative-convective equilibrium profile, following the description set out
139     in Held and Suarez; 1994 designed to test atmospheric hydrodynamical cores -
140     there are no mountains or land-sea contrast.
141    
142     As described in Adcroft (2001), a `cubed sphere' is used to discretize the
143     globe permitting a uniform gridding and obviated the need to fourier filter.
144     The `vector-invariant' form of MITgcm supports any orthogonal curvilinear
145     grid, of which the cubed sphere is just one of many choices.
146 cnh 1.1
147     Fig.E1b shows the 5-year mean, zonally averaged potential temperature, zonal
148 cnh 1.2 wind and meridional overturning streamfunction from a 20-level version of
149     the model. It compares favorable with more conventional spatial
150     discretization approaches.
151 cnh 1.1
152     A regular spherical lat-lon grid can also be used.
153    
154 cnh 1.2 \subsection{Ocean gyres}
155 cnh 1.1
156 cnh 1.2 Baroclinic instability is a ubiquitous process in the ocean, as well as the
157     atmosphere. Ocean eddies play an important role in modifying the
158     hydrographic structure and current systems of the oceans. Coarse resolution
159     models of the oceans cannot resolve the eddy field and yield rather broad,
160     diffusive patterns of ocean currents. But if the resolution of our models is
161     increased until the baroclinic instability process is resolved, numerical
162     solutions of a different and much more realistic kind, can be obtained.
163    
164     Fig. ?.? shows the surface temperature and velocity field obtained from
165     MITgcm run at $\frac{1}{6}^{\circ }$ horizontal resolution on a $lat-lon$
166     grid in which the pole has been rotated by 90$^{\circ }$ on to the equator
167     (to avoid the converging of meridian in northern latitudes). 21 vertical
168     levels are used in the vertical with a `lopped cell' representation of
169     topography. The development and propagation of anomalously warm and cold
170     eddies can be clearly been seen in the Gulf Stream region. The transport of
171     warm water northward by the mean flow of the Gulf Stream is also clearly
172     visible.
173 cnh 1.1
174     \subsection{Global ocean circulation}
175    
176     Fig.E2a shows the pattern of ocean currents at the surface of a 4$^{\circ }$
177 cnh 1.2 global ocean model run with 15 vertical levels. Lopped cells are used to
178     represent topography on a regular $lat-lon$ grid extending from 70$^{\circ
179     }N $ to 70$^{\circ }S$. The model is driven using monthly-mean winds with
180     mixed boundary conditions on temperature and salinity at the surface. The
181     transfer properties of ocean eddies, convection and mixing is parameterized
182     in this model.
183    
184     Fig.E2b shows the meridional overturning circulation of the global ocean in
185     Sverdrups.
186    
187     \subsection{Convection and mixing over topography}
188    
189     Dense plumes generated by localized cooling on the continental shelf of the
190     ocean may be influenced by rotation when the deformation radius is smaller
191     than the width of the cooling region. Rather than gravity plumes, the
192     mechanism for moving dense fluid down the shelf is then through geostrophic
193     eddies. The simulation shown in the figure (blue is cold dense fluid, red is
194     warmer, lighter fluid) employs the non-hydrostatic capability of MITgcm to
195     trigger convection by surface cooling. The cold, dense water falls down the
196     slope but is deflected along the slope by rotation. It is found that
197     entrainment in the vertical plane is reduced when rotational control is
198     strong, and replaced by lateral entrainment due to the baroclinic
199     instability of the along-slope current.
200 cnh 1.1
201     \subsection{Boundary forced internal waves}
202    
203 cnh 1.2 The unique ability of MITgcm to treat non-hydrostatic dynamics in the
204     presence of complex geometry makes it an ideal tool to study internal wave
205     dynamics and mixing in oceanic canyons and ridges driven by large amplitude
206     barotropic tidal currents imposed through open boundary conditions.
207    
208     Fig. ?.? shows the influence of cross-slope topographic variations on
209     internal wave breaking - the cross-slope velocity is in color, the density
210     contoured. The internal waves are excited by application of open boundary
211     conditions on the left.\ They propagate to the sloping boundary (represented
212     using MITgcm's finite volume spatial discretization) where they break under
213     nonhydrostatic dynamics.
214    
215     \subsection{Parameter sensitivity using the adjoint of MITgcm}
216    
217     Forward and tangent linear counterparts of MITgcm are supported using an
218     `automatic adjoint compiler'. These can be used in parameter sensitivity and
219     data assimilation studies.
220    
221     As one example of application of the MITgcm adjoint, Fig.E4 maps the
222     gradient $\frac{\partial J}{\partial \mathcal{H}}$where $J$ is the magnitude
223     of the overturning streamfunction shown in fig?.? at 40$^{\circ }$N and $%
224     \mathcal{H}$ is the air-sea heat flux 100 years before. We see that $J$ is
225     sensitive to heat fluxes over the Labrador Sea, one of the important sources
226     of deep water for the thermohaline circulations. This calculation also
227     yields sensitivities to all other model parameters.
228    
229     \subsection{Global state estimation of the ocean}
230    
231     An important application of MITgcm is in state estimation of the global
232     ocean circulation. An appropriately defined `cost function', which measures
233     the departure of the model from observations (both remotely sensed and
234     insitu) over an interval of time, is minimized by adjusting `control
235     parameters' such as air-sea fluxes, the wind field, the initial conditions
236     etc. Figure ?.? shows an estimate of the time-mean surface elevation of the
237     ocean obtained by bringing the model in to consistency with altimetric and
238     in-situ observations over the period 1992-1997.
239    
240     \subsection{Ocean biogeochemical cycles}
241    
242     MITgcm is being used to study global biogeochemical cycles in the ocean. For
243     example one can study the effects of interannual changes in meteorological
244     forcing and upper ocean circulation on the fluxes of carbon dioxide and
245     oxygen between the ocean and atmosphere. The figure shows the annual air-sea
246     flux of oxygen and its relation to density outcrops in the southern oceans
247     from a single year of a global, interannually varying simulation.
248    
249     Chris - get figure here: http://puddle.mit.edu/\symbol{126}%
250     mick/biogeochem.html
251    
252     \subsection{Simulations of laboratory experiments}
253    
254     Figure ?.? shows MITgcm being used to simulate a laboratory experiment
255     enquiring in to the dynamics of the Antarctic Circumpolar Current (ACC). An
256     initially homogeneous tank of water ($1m$ in diameter) is driven from its
257     free surface by a rotating heated disk. The combined action of mechanical
258     and thermal forcing creates a lens of fluid which becomes baroclinically
259     unstable. The stratification and depth of penetration of the lens is
260     arrested by its instability in a process analogous to that whic sets the
261     stratification of the ACC.
262 cnh 1.1
263     % $Header: /u/gcmpack/mitgcmdoc/part1/continuous_eqns.tex,v 1.3 2001/09/26 14:53:10 cnh Exp $
264     % $Name: $
265    
266     \section{Continuous equations in `r' coordinates}
267    
268     To render atmosphere and ocean models from one dynamical core we exploit
269     `isomorphisms' between equation sets that govern the evolution of the
270     respective fluids - see fig.4%
271     \marginpar{
272     Fig.4. Isomorphisms}. One system of hydrodynamical equations is written down
273     and encoded. The model variables have different interpretations depending on
274     whether the atmosphere or ocean is being studied. Thus, for example, the
275     vertical coordinate `$r$' is interpreted as pressure, $p$, if we are
276     modeling the atmosphere and height, $z$, if we are modeling the ocean.
277    
278     The state of the fluid at any time is characterized by the distribution of
279     velocity $\vec{\mathbf{v}}$, active tracers $\theta $ and $S$, a
280     `geopotential' $\phi $ and density $\rho =\rho (\theta ,S,p)$ which may
281     depend on $\theta $, $S$, and $p$. The equations that govern the evolution
282     of these fields, obtained by applying the laws of classical mechanics and
283     thermodynamics to a Boussinesq, Navier-Stokes fluid are, written in terms of
284     a generic vertical coordinate, $r$, see fig.5%
285     \marginpar{
286     Fig.5 The vertical coordinate of model}:
287    
288     \begin{equation*}
289     \frac{D\vec{\mathbf{v}_{h}}}{Dt}+\left( 2\vec{\Omega}\times \vec{\mathbf{v}}%
290     \right) _{h}+\mathbf{\nabla }_{h}\phi =\mathcal{F}_{\vec{\mathbf{v}_{h}}}%
291     \text{ horizontal mtm}
292     \end{equation*}
293    
294     \begin{equation*}
295     \frac{D\dot{r}}{Dt}+\widehat{k}\cdot \left( 2\vec{\Omega}\times \vec{\mathbf{%
296     v}}\right) +\frac{\partial \phi }{\partial r}+b=\mathcal{F}_{\dot{r}}\text{
297     vertical mtm}
298     \end{equation*}
299    
300     \begin{equation}
301     \mathbf{\nabla }_{h}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \dot{r}}{%
302     \partial r}=0\text{ continuity} \label{eq:continuous}
303     \end{equation}
304    
305     \begin{equation*}
306 cnh 1.2 b=b(\theta ,S,r)\text{ equation of state}
307 cnh 1.1 \end{equation*}
308    
309     \begin{equation*}
310 cnh 1.2 \frac{D\theta }{Dt}=\mathcal{Q}_{\theta }\text{ potential temperature}
311 cnh 1.1 \end{equation*}
312    
313     \begin{equation*}
314 cnh 1.2 \frac{DS}{Dt}=\mathcal{Q}_{S}\text{ humidity/salinity}
315 cnh 1.1 \end{equation*}
316    
317     Here:
318    
319     \begin{equation*}
320 cnh 1.2 r\text{ is the vertical coordinate}
321 cnh 1.1 \end{equation*}
322    
323     \begin{equation*}
324     \frac{D}{Dt}=\frac{\partial }{\partial t}+\vec{\mathbf{v}}\cdot \nabla \text{
325 cnh 1.2 is the total derivative}
326 cnh 1.1 \end{equation*}
327    
328     \begin{equation*}
329     \mathbf{\nabla }=\mathbf{\nabla }_{h}+\widehat{k}\frac{\partial }{\partial r}%
330 cnh 1.2 \text{ is the `grad' operator}
331 cnh 1.1 \end{equation*}
332     with $\mathbf{\nabla }_{h}$ operating in the horizontal and $\widehat{k}%
333     \frac{\partial }{\partial r}$ operating in the vertical, where $\widehat{k}$
334     is a unit vector in the vertical
335    
336     \begin{equation*}
337 cnh 1.2 t\text{ is time}
338 cnh 1.1 \end{equation*}
339    
340     \begin{equation*}
341     \vec{\mathbf{v}}=(u,v,\dot{r})=(\vec{\mathbf{v}}_{h},\dot{r})\text{ is the
342 cnh 1.2 velocity}
343 cnh 1.1 \end{equation*}
344    
345     \begin{equation*}
346 cnh 1.2 \phi \text{ is the `pressure'/`geopotential'}
347 cnh 1.1 \end{equation*}
348    
349     \begin{equation*}
350 cnh 1.2 \vec{\Omega}\text{ is the Earth's rotation}
351 cnh 1.1 \end{equation*}
352    
353     \begin{equation*}
354 cnh 1.2 b\text{ is the `buoyancy'}
355 cnh 1.1 \end{equation*}
356    
357     \begin{equation*}
358 cnh 1.2 \theta \text{ is potential temperature}
359 cnh 1.1 \end{equation*}
360    
361     \begin{equation*}
362 cnh 1.2 S\text{ is specific humidity in the atmosphere; salinity in the ocean}
363 cnh 1.1 \end{equation*}
364    
365     \begin{equation*}
366     \mathcal{F}_{\vec{\mathbf{v}}}\text{ are forcing and dissipation of }\vec{%
367     \mathbf{v}}
368     \end{equation*}
369    
370     \begin{equation*}
371 cnh 1.2 \mathcal{Q}_{\theta }\mathcal{\ }\text{are forcing and dissipation of }\theta
372 cnh 1.1 \end{equation*}
373    
374     \begin{equation*}
375     \mathcal{Q}_{S}\mathcal{\ }\text{are forcing and dissipation of }S
376     \end{equation*}
377    
378     The $\mathcal{F}^{\prime }s$ and $\mathcal{Q}^{\prime }s$ are provided by
379     extensive `physics' packages for atmosphere and ocean described in Chapter 6.
380    
381     \subsection{Kinematic Boundary conditions}
382    
383     \subsubsection{vertical}
384    
385     at fixed and moving $r$ surfaces we set (see fig.5):
386    
387     \begin{equation}
388     \dot{r}=0atr=R_{fixed}(x,y)\text{ (ocean bottom, top of the atmosphere)}
389     \label{eq:fixedbc}
390     \end{equation}
391    
392     \begin{equation}
393     \dot{r}=\frac{Dr}{Dt}atr=R_{moving}\text{ \
394     (oceansurface,bottomoftheatmosphere)} \label{eq:movingbc}
395     \end{equation}
396    
397     Here
398    
399     \begin{equation*}
400 cnh 1.2 R_{moving}=R_{o}+\eta
401 cnh 1.1 \end{equation*}
402     where $R_{o}(x,y)$ is the `$r-$value' (height or pressure, depending on
403     whether we are in the atmosphere or ocean) of the `moving surface' in the
404     resting fluid and $\eta $ is the departure from $R_{o}(x,y)$ in the presence
405     of motion.
406    
407     \subsubsection{horizontal}
408    
409     \begin{equation}
410     \vec{\mathbf{v}}\cdot \vec{\mathbf{n}}=0 \label{eq:noflow}
411     \end{equation}%
412     where $\vec{\mathbf{n}}$ is the normal to a solid boundary.
413    
414     \subsection{Atmosphere}
415    
416     In the atmosphere, see fig.5, we interpret:
417    
418     \begin{equation}
419     r=p\text{ is the pressure} \label{eq:atmos-r}
420     \end{equation}
421    
422     \begin{equation}
423     \dot{r}=\frac{Dp}{Dt}=\omega \text{ is the vertical velocity in }p\text{
424     coordinates} \label{eq:atmos-omega}
425     \end{equation}
426    
427     \begin{equation}
428     \phi =g\,z\text{ is the geopotential height} \label{eq:atmos-phi}
429     \end{equation}
430    
431     \begin{equation}
432     b=\frac{\partial \Pi }{\partial p}\theta \text{ is the buoyancy}
433     \label{eq:atmos-b}
434     \end{equation}
435    
436     \begin{equation}
437     \theta =T(\frac{p_{c}}{p})^{\kappa }\text{ is potential temperature}
438     \label{eq:atmos-theta}
439     \end{equation}
440    
441     \begin{equation}
442     S=q,\text{ is the specific humidity} \label{eq:atmos-s}
443     \end{equation}
444     where
445    
446     \begin{equation*}
447     T\text{ is absolute temperature}
448     \end{equation*}%
449     \begin{equation*}
450     p\text{ is the pressure}
451     \end{equation*}%
452     \begin{eqnarray*}
453     &&z\text{ is the height of the pressure surface} \\
454     &&g\text{ is the acceleration due to gravity}
455     \end{eqnarray*}
456    
457     In the above the ideal gas law, $p=\rho RT$, has been expressed in terms of
458     the Exner function $\Pi (p)$ given by (see Appendix Atmosphere)
459     \begin{equation}
460     \Pi (p)=c_{p}(\frac{p}{p_{c}})^{\kappa } \label{eq:exner}
461     \end{equation}%
462     where $p_{c}$ is a reference pressure and $\kappa =R/c_{p}$ with $R$ the gas
463     constant and $c_{p}$ the specific heat of air at constant pressure.
464    
465     At the top of the atmosphere (which is `fixed' in our $r$ coordinate):
466    
467     \begin{equation*}
468 cnh 1.2 R_{fixed}=p_{top}=0
469 cnh 1.1 \end{equation*}
470     In a resting atmosphere the elevation of the mountains at the bottom is
471     given by
472     \begin{equation*}
473 cnh 1.2 R_{moving}=R_{o}(x,y)=p_{o}(x,y)
474 cnh 1.1 \end{equation*}
475     i.e. the (hydrostatic) pressure at the top of the mountains in a resting
476     atmosphere.
477    
478     The boundary conditions at top and bottom are given by:
479    
480     \begin{eqnarray}
481     &&\omega =0~\text{at }r=R_{fixed} \text{ (top of the atmosphere)}
482     \label{eq:fixed-bc-atmos} \\
483     \omega &=&\frac{Dp_{s}}{Dt}\text{; at }r=R_{moving}\text{ (bottom of the
484     atmosphere)} \label{eq:moving-bc-atmos}
485     \end{eqnarray}
486    
487     Then the (hydrostatic form of) eq(\ref{eq:continuous}) yields a consistent
488     set of atmospheric equations which, for convenience, are written out in $p$
489     coordinates in Appendix Atmosphere - see eqs(\ref{eq:atmos-prime}).
490    
491     \subsection{Ocean}
492    
493     In the ocean we interpret:
494     \begin{eqnarray}
495     r &=&z\text{ is the height} \label{eq:ocean-z} \\
496     \dot{r} &=&\frac{Dz}{Dt}=w\text{ is the vertical velocity}
497     \label{eq:ocean-w} \\
498     \phi &=&\frac{p}{\rho _{c}}\text{ is the pressure} \label{eq:ocean-p} \\
499     b(\theta ,S,r) &=&\frac{g}{\rho _{c}}\left( \rho (\theta ,S,r)-\rho
500     _{c}\right) \text{ is the buoyancy} \label{eq:ocean-b}
501     \end{eqnarray}
502     where $\rho _{c}$ is a fixed reference density of water and $g$ is the
503     acceleration due to gravity.\noindent
504    
505     In the above
506    
507     At the bottom of the ocean: $R_{fixed}(x,y)=-H(x,y)$.
508    
509     The surface of the ocean is given by: $R_{moving}=\eta $
510    
511     The position of the resting free surface of the ocean is given by $%
512     R_{o}=Z_{o}=0$.
513    
514     Boundary conditions are:
515    
516     \begin{eqnarray}
517     w &=&0~\text{at }r=R_{fixed}\text{ (ocean bottom)} \label{eq:fixed-bc-ocean}
518     \\
519     w &=&\frac{D\eta }{Dt}\text{ at }r=R_{moving}=\eta \text{ (ocean surface) %
520     \label{eq:moving-bc-ocean}}
521     \end{eqnarray}
522     where $\eta $ is the elevation of the free surface.
523    
524     Then eq(\ref{eq:continuous}) yields a consistent set of oceanic equations
525     which, for convenience, are written out in $z$ coordinates in Appendix Ocean
526     - see eqs(\ref{eq:ocean-mom}) to (\ref{eq:ocean-salt}).
527    
528     \subsection{Hydrostatic, Quasi-hydrostatic, Quasi-nonhydrostatic and
529     Non-hydrostatic forms}
530    
531     Let us separate $\phi $ in to surface, hydrostatic and non-hydrostatic terms:
532    
533     \begin{equation}
534     \phi (x,y,r)=\phi _{s}(x,y)+\phi _{hyd}(x,y,r)+\phi _{nh}(x,y,r)
535     \label{eq:phi-split}
536     \end{equation}%
537     and write eq(\ref{incompressible}a,b) in the form:
538    
539     \begin{equation}
540     \frac{\partial \vec{\mathbf{v}_{h}}}{\partial t}+\mathbf{\nabla }_{h}\phi
541     _{s}+\mathbf{\nabla }_{h}\phi _{hyd}+\epsilon _{nh}\mathbf{\nabla }_{h}\phi
542     _{nh}=\vec{\mathbf{G}}_{\vec{v}_{h}} \label{eq:mom-h}
543     \end{equation}
544    
545     \begin{equation}
546     \frac{\partial \phi _{hyd}}{\partial r}=-b \label{eq:hydrostatic}
547     \end{equation}
548    
549     \begin{equation}
550     \epsilon _{nh}\frac{\partial \dot{r}}{\partial t}+\frac{\partial \phi _{nh}}{%
551     \partial r}=G_{\dot{r}} \label{eq:mom-w}
552     \end{equation}
553     Here $\epsilon _{nh}$ is a non-hydrostatic parameter.
554    
555     The $\left( \vec{\mathbf{G}}_{\vec{v}},G_{\dot{r}}\right) $ in eq(\ref%
556     {eq:mom-h}) and (\ref{eq:mom-w}) represent advective, metric and Coriolis
557     terms in the momentum equations. In spherical coordinates they take the form%
558     \footnote{%
559     In the hydrostatic primitive equations (\textbf{HPE}) all underlined terms
560     in (\ref{eq:gu-speherical}), (\ref{eq:gv-spherical}) and (\ref%
561     {eq:gw-spherical}) are omitted; the singly-underlined terms are included in
562     the quasi-hydrostatic model (\textbf{QH}). The fully non-hydrostatic model (%
563     \textbf{NH}) includes all terms.} - see Marshall et al 1997a for a full
564     discussion:
565    
566     \begin{equation}
567     \left.
568     \begin{tabular}{l}
569     $G_{u}=-\vec{\mathbf{v}}.\nabla u$ \\
570     $-\left\{ \underline{\frac{u\dot{r}}{{r}}}-\frac{uv\tan lat}{{r}}\right\} $
571     \\
572     $-\left\{ -2\Omega v\sin lat+\underline{2\Omega \dot{r}\cos lat}\right\} $
573     \\
574     $+\mathcal{F}_{u}$%
575     \end{tabular}%
576     \ \right\} \left\{
577     \begin{tabular}{l}
578     \textit{advection} \\
579     \textit{metric} \\
580     \textit{Coriolis} \\
581     \textit{\ Forcing/Dissipation}%
582     \end{tabular}%
583 cnh 1.2 \ \right. \qquad \label{eq:gu-speherical}
584 cnh 1.1 \end{equation}
585    
586     \begin{equation}
587     \left.
588     \begin{tabular}{l}
589     $G_{v}=-\vec{\mathbf{v}}.\nabla v$ \\
590     $-\left\{ \underline{\frac{v\dot{r}}{{r}}}-\frac{u^{2}\tan lat}{{r}}\right\}
591     $ \\
592     $-\left\{ -2\Omega u\sin lat\right\} $ \\
593     $+\mathcal{F}_{v}$%
594     \end{tabular}%
595     \ \right\} \left\{
596     \begin{tabular}{l}
597     \textit{advection} \\
598     \textit{metric} \\
599     \textit{Coriolis} \\
600     \textit{\ Forcing/Dissipation}%
601     \end{tabular}%
602 cnh 1.2 \ \right. \qquad \label{eq:gv-spherical}
603 cnh 1.1 \end{equation}%
604 cnh 1.2 \qquad \qquad \qquad \qquad \qquad
605 cnh 1.1
606     \begin{equation}
607     \left.
608     \begin{tabular}{l}
609     $G_{\dot{r}}=-\underline{\underline{\vec{\mathbf{v}}.\nabla \dot{r}}}$ \\
610     $+\left\{ \underline{\frac{u^{_{^{2}}}+v^{2}}{{r}}}\right\} $ \\
611     ${+}\underline{{2\Omega u\cos lat}}$ \\
612     $\underline{\underline{\mathcal{F}_{\dot{r}}}}$%
613     \end{tabular}%
614     \ \right\} \left\{
615     \begin{tabular}{l}
616     \textit{advection} \\
617     \textit{metric} \\
618     \textit{Coriolis} \\
619     \textit{\ Forcing/Dissipation}%
620     \end{tabular}%
621 cnh 1.2 \ \right. \label{eq:gw-spherical}
622 cnh 1.1 \end{equation}%
623 cnh 1.2 \qquad \qquad \qquad \qquad \qquad
624 cnh 1.1
625     In the above `${r}$' is the distance from the center of the earth and `$lat$%
626     ' is latitude.
627    
628     Grad and div operators in spherical coordinates are defined in appendix
629     OPERATORS.%
630     \marginpar{
631     Fig.6 Spherical polar coordinate system.}
632    
633     \subsubsection{Shallow atmosphere approximation}
634    
635     Most models are based on the `hydrostatic primitive equations' (HPE's) in
636     which the vertical momentum equation is reduced to a statement of
637     hydrostatic balance and the `traditional approximation' is made in which the
638     Coriolis force is treated approximately and the shallow atmosphere
639     approximation is made.\ The MITgcm need not make the `traditional
640     approximation'. To be able to support consistent non-hydrostatic forms the
641 cnh 1.2 shallow atmosphere approximation can be relaxed - when dividing through by $%
642     r $ in, for example, (\ref{eq:gu-speherical}), we do not replace $r$ by $a$,
643 cnh 1.1 the radius of the earth.
644    
645     \subsubsection{Hydrostatic and quasi-hydrostatic forms}
646    
647     These are discussed at length in Marshall et al (1997a).
648    
649     In the `hydrostatic primitive equations' (\textbf{HPE)} all the underlined
650     terms in Eqs. (\ref{eq:gu-speherical} $\rightarrow $\ \ref{eq:gw-spherical})
651     are neglected and `${r}$' is replaced by `$a$', the mean radius of the
652     earth. Once the pressure is found at one level - e.g. by inverting a 2-d
653     Elliptic equation for $\phi _{s}$ at $r=R_{moving}$ - the pressure can be
654     computed at all other levels by integration of the hydrostatic relation, eq(%
655     \ref{eq:hydrostatic}).
656    
657     In the `quasi-hydrostatic' equations (\textbf{QH)} strict balance between
658     gravity and vertical pressure gradients is not imposed. The $2\Omega u\cos
659     \phi $ Coriolis term are not neglected and are balanced by a non-hydrostatic
660     contribution to the pressure field: only the terms underlined twice in Eqs. (%
661     \ref{eq:gu-speherical}$\rightarrow $\ \ref{eq:gw-spherical}) are set to zero
662     and, simultaneously, the shallow atmosphere approximation is relaxed. In
663     \textbf{QH}\ \textit{all} the metric terms are retained and the full
664     variation of the radial position of a particle monitored. The \textbf{QH}\
665     vertical momentum equation (\ref{eq:mom-w}) becomes:
666    
667     \begin{equation*}
668 cnh 1.2 \frac{\partial \phi _{nh}}{\partial r}=2\Omega u\cos lat
669 cnh 1.1 \end{equation*}
670     making a small correction to the hydrostatic pressure.
671    
672     \textbf{QH} has good energetic credentials - they are the same as for
673     \textbf{HPE}. Importantly, however, it has the same angular momentum
674     principle as the full non-hydrostatic model (\textbf{NH)} - see Marshall
675     et.al., 1997a. As in \textbf{HPE }only a 2-d elliptic problem need be solved.
676    
677     \subsubsection{Non-hydrostatic and quasi-nonhydrostatic forms}
678    
679     The MIT model presently supports a full non-hydrostatic ocean isomorph, but
680     only a quasi-non-hydrostatic atmospheric isomorph.
681    
682     \paragraph{Non-hydrostatic Ocean}
683    
684     In the non-hydrostatic ocean model all terms in equations Eqs.(\ref%
685     {eq:gu-speherical} $\rightarrow $\ \ref{eq:gw-spherical}) are retained. A
686     three dimensional elliptic equation must be solved subject to Neumann
687     boundary conditions (see below). It is important to note that use of the
688     full \textbf{NH} does not admit any new `fast' waves in to the system - the
689     incompressible condition eq(\ref{eq:continuous})c has already filtered out
690     acoustic modes. It does, however, ensure that the gravity waves are treated
691     accurately with an exact dispersion relation. The \textbf{NH} set has a
692     complete angular momentum principle and consistent energetics - see White
693     and Bromley, 1995; Marshall et.al.\ 1997a.
694    
695     \paragraph{Quasi-nonhydrostatic Atmosphere}
696    
697     In the non-hydrostatic version of our atmospheric model we approximate $\dot{%
698     r}$ in the vertical momentum eqs(\ref{eq:mom-w}) and (\ref{eq:gv-spherical})
699     (but only here) by:
700    
701     \begin{equation}
702     \dot{r}=\frac{Dp}{Dt}=\frac{1}{g}\frac{D\phi }{Dt} \label{eq:quasi-nh-w}
703     \end{equation}%
704     where $p_{hy}$ is the hydrostatic pressure.
705    
706     \subsubsection{Summary of equation sets supported by model}
707    
708     \paragraph{Atmosphere}
709    
710     Hydrostatic, and quasi-hydrostatic and quasi non-hydrostatic forms of the
711     compressible non-Boussinesq equations in $p-$coordinates are supported.
712    
713     \subparagraph{Hydrostatic and quasi-hydrostatic}
714    
715     The hydrostatic set is written out in $p-$coordinates in appendix Atmosphere
716     - see eq(\ref{eq:atmos-prime}).
717    
718     \subparagraph{Quasi-nonhydrostatic}
719    
720     A quasi-nonhydrostatic form is also supported.
721    
722     \paragraph{Ocean}
723    
724     \subparagraph{Hydrostatic and quasi-hydrostatic}
725    
726     Hydrostatic, and quasi-hydrostatic forms of the incompressible Boussinesq
727     equations in $z-$coordinates are supported.
728    
729     \subparagraph{Non-hydrostatic}
730    
731     Non-hydrostatic forms of the incompressible Boussinesq equations in $z-$%
732     coordinates are supported - see eqs(\ref{eq:ocean-mom}) to (\ref%
733     {eq:ocean-salt}).
734    
735     \subsection{Solution strategy}
736    
737     The method of solution employed in the \textbf{HPE}, \textbf{QH} and \textbf{%
738     NH} models is summarized in Fig.7.%
739     \marginpar{
740     Fig.7 Solution strategy} Under all dynamics, a 2-d elliptic equation is
741     first solved to find the surface pressure and the hydrostatic pressure at
742     any level computed from the weight of fluid above. Under \textbf{HPE} and
743     \textbf{QH} dynamics, the horizontal momentum equations are then stepped
744     forward and $\dot{r}$ found from continuity. Under \textbf{NH} dynamics a
745     3-d elliptic equation must be solved for the non-hydrostatic pressure before
746     stepping forward the horizontal momentum equations; $\dot{r}$ is found by
747     stepping forward the vertical momentum equation.
748    
749     There is no penalty in implementing \textbf{QH} over \textbf{HPE} except, of
750     course, some complication that goes with the inclusion of $\cos \phi \ $%
751     Coriolis terms and the relaxation of the shallow atmosphere approximation.
752     But this leads to negligible increase in computation. In \textbf{NH}, in
753     contrast, one additional elliptic equation - a three-dimensional one - must
754     be inverted for $p_{nh}$. However the `overhead' of the \textbf{NH} model is
755     essentially negligible in the hydrostatic limit (see detailed discussion in
756     Marshall et al, 1997) resulting in a non-hydrostatic algorithm that, in the
757     hydrostatic limit, is as computationally economic as the \textbf{HPEs}.
758    
759     \subsection{Finding the pressure field}
760    
761     Unlike the prognostic variables $u$, $v$, $w$, $\theta $ and $S$, the
762     pressure field must be obtained diagnostically. We proceed, as before, by
763     dividing the total (pressure/geo) potential in to three parts, a surface
764     part, $\phi _{s}(x,y)$, a hydrostatic part $\phi _{hyd}(x,y,r)$ and a
765     non-hydrostatic part $\phi _{nh}(x,y,r)$, as in (\ref{eq:phi-split}), and
766     writing the momentum equation as in (\ref{eq:mom-h}).
767    
768     \subsubsection{Hydrostatic pressure}
769    
770     Hydrostatic pressure is obtained by integrating (\ref{eq:hydrostatic})
771     vertically from $r=R_{o}$ where $\phi _{hyd}(r=R_{o})=0$, to yield:
772    
773     \begin{equation*}
774     \int_{r}^{R_{o}}\frac{\partial \phi _{hyd}}{\partial r}dr=\left[ \phi _{hyd}%
775 cnh 1.2 \right] _{r}^{R_{o}}=\int_{r}^{R_{o}}-bdr
776 cnh 1.1 \end{equation*}
777     and so
778    
779     \begin{equation}
780     \phi _{hyd}(x,y,r)=\int_{r}^{R_{o}}bdr \label{eq:hydro-phi}
781     \end{equation}
782    
783     The model can be easily modified to accommodate a loading term (e.g
784     atmospheric pressure pushing down on the ocean's surface) by setting:
785    
786     \begin{equation}
787     \phi _{hyd}(r=R_{o})=loading \label{eq:loading}
788     \end{equation}
789    
790     \subsubsection{Surface pressure}
791    
792     The surface pressure equation can be obtained by integrating continuity, (%
793     \ref{eq:continuous})c, vertically from $r=R_{fixed}$ to $r=R_{moving}$
794    
795     \begin{equation*}
796     \int_{R_{fixed}}^{R_{moving}}\left( \mathbf{\nabla }_{h}\cdot \vec{\mathbf{v}%
797 cnh 1.2 }_{h}+\partial _{r}\dot{r}\right) dr=0
798 cnh 1.1 \end{equation*}
799    
800     Thus:
801    
802     \begin{equation*}
803     \frac{\partial \eta }{\partial t}+\vec{\mathbf{v}}.\nabla \eta
804     +\int_{R_{fixed}}^{R_{moving}}\mathbf{\nabla }_{h}\cdot \vec{\mathbf{v}}%
805 cnh 1.2 _{h}dr=0
806 cnh 1.1 \end{equation*}
807     where $\eta =R_{moving}-R_{o}$ is the free-surface $r$-anomaly in units of $%
808     r $. The above can be rearranged to yield, using Leibnitz's theorem:
809    
810     \begin{equation}
811     \frac{\partial \eta }{\partial t}+\mathbf{\nabla }_{h}\cdot
812     \int_{R_{fixed}}^{R_{moving}}\vec{\mathbf{v}}_{h}dr=\text{source}
813     \label{eq:free-surface}
814     \end{equation}%
815     where we have incorporated a source term.
816    
817     Whether $\phi $ is pressure (ocean model, $p/\rho _{c}$) or geopotential
818     (atmospheric model), in (\ref{mtm-split}), the horizontal gradient term can
819     be written
820     \begin{equation}
821 cnh 1.2 \mathbf{\nabla }_{h}\phi _{s}=\mathbf{\nabla }_{h}\left( b_{s}\eta \right)
822 cnh 1.1 \label{eq:phi-surf}
823     \end{equation}%
824     where $b_{s}$ is the buoyancy at the surface.
825    
826     In the hydrostatic limit ($\epsilon _{nh}=0$), Eqs(\ref{eq:mom-h}), (\ref%
827     {eq:free-surface}) and (\ref{eq:phi-surf}) can be solved by inverting a 2-d
828     elliptic equation for $\phi _{s}$ as described in Chapter 2. Both `free
829     surface' and `rigid lid' approaches are available.
830    
831     \subsubsection{Non-hydrostatic pressure}
832    
833     Taking the horizontal divergence of (\ref{hor-mtm}) and adding $\frac{%
834     \partial }{\partial r}$ of (\ref{vertmtm}), invoking the continuity equation
835     (\ref{incompressible}), we deduce that:
836    
837     \begin{equation}
838     \nabla _{3}^{2}\phi _{nh}=\nabla .\vec{\mathbf{G}}_{\vec{v}}-\left( \mathbf{%
839     \nabla }_{h}^{2}\phi _{s}+\mathbf{\nabla }^{2}\phi _{hyd}\right) =\nabla .%
840     \vec{\mathbf{F}} \label{eq:3d-invert}
841     \end{equation}
842    
843     For a given rhs this 3-d elliptic equation must be inverted for $\phi _{nh}$
844     subject to appropriate choice of boundary conditions. This method is usually
845     called \textit{The Pressure Method} [Harlow and Welch, 1965; Williams, 1969;
846     Potter, 1976]. In the hydrostatic primitive equations case (\textbf{HPE}),
847     the 3-d problem does not need to be solved.
848    
849     \paragraph{Boundary Conditions}
850    
851     We apply the condition of no normal flow through all solid boundaries - the
852     coasts (in the ocean) and the bottom:
853    
854     \begin{equation}
855     \vec{\mathbf{v}}.\widehat{n}=0 \label{nonormalflow}
856     \end{equation}
857     where $\widehat{n}$ is a vector of unit length normal to the boundary. The
858     kinematic condition (\ref{nonormalflow}) is also applied to the vertical
859     velocity at $r=R_{moving}$. No-slip $\left( v_{T}=0\right) \ $or slip $%
860     \left( \partial v_{T}/\partial n=0\right) \ $conditions are employed on the
861     tangential component of velocity, $v_{T}$, at all solid boundaries,
862     depending on the form chosen for the dissipative terms in the momentum
863     equations - see below.
864    
865     Eq.(\ref{nonormalflow}) implies, making use of (\ref{mtm-split}), that:
866    
867     \begin{equation}
868     \widehat{n}.\nabla \phi _{nh}=\widehat{n}.\vec{\mathbf{F}}
869     \label{eq:inhom-neumann-nh}
870     \end{equation}
871     where
872    
873     \begin{equation*}
874     \vec{\mathbf{F}}=\vec{\mathbf{G}}_{\vec{v}}-\left( \mathbf{\nabla }_{h}\phi
875     _{s}+\mathbf{\nabla }\phi _{hyd}\right)
876     \end{equation*}%
877     presenting inhomogeneous Neumann boundary conditions to the Elliptic problem
878     (\ref{eq:3d-invert}). As shown, for example, by Williams (1969), one can
879     exploit classical 3D potential theory and, by introducing an appropriately
880 cnh 1.2 chosen $\delta $-function sheet of `source-charge', replace the
881     inhomogeneous boundary condition on pressure by a homogeneous one. The
882     source term $rhs$ in (\ref{eq:3d-invert}) is the divergence of the vector $%
883     \vec{\mathbf{F}}.$ By simultaneously setting $%
884 cnh 1.1 \begin{array}{l}
885     \widehat{n}.\vec{\mathbf{F}}%
886     \end{array}%
887     =0$\ and $\widehat{n}.\nabla \phi _{nh}=0\ $on the boundary the following
888 cnh 1.2 self-consistent but simpler homogenized Elliptic problem is obtained:
889 cnh 1.1
890     \begin{equation*}
891 cnh 1.2 \nabla ^{2}\phi _{nh}=\nabla .\widetilde{\vec{\mathbf{F}}}\qquad
892 cnh 1.1 \end{equation*}%
893     where $\widetilde{\vec{\mathbf{F}}}$ is a modified $\vec{\mathbf{F}}$ such
894     that $\widetilde{\vec{\mathbf{F}}}.\widehat{n}=0$. As is implied by (\ref%
895     {eq:inhom-neumann-nh}) the modified boundary condition becomes:
896    
897     \begin{equation}
898     \widehat{n}.\nabla \phi _{nh}=0 \label{eq:hom-neumann-nh}
899     \end{equation}
900    
901     If the flow is `close' to hydrostatic balance then the 3-d inversion
902     converges rapidly because $\phi _{nh}\ $is then only a small correction to
903     the hydrostatic pressure field (see the discussion in Marshall et al, a,b).
904    
905     The solution $\phi _{nh}\ $to (\ref{eq:3d-invert}) and (\ref{homneuman})
906     does not vanish at $r=R_{moving}$, and so refines the pressure there.
907    
908     \subsection{Forcing/dissipation}
909    
910     \subsubsection{Forcing}
911    
912     The forcing terms $\mathcal{F}$ on the rhs of the equations are provided by
913     `physics packages' described in detail in chapter ??.
914    
915     \subsubsection{Dissipation}
916    
917     \paragraph{Momentum}
918    
919     Many forms of momentum dissipation are available in the model. Laplacian and
920     biharmonic frictions are commonly used:
921    
922     \begin{equation}
923     D_{V}=A_{h}\nabla _{h}^{2}v+A_{v}\frac{\partial ^{2}v}{\partial z^{2}}%
924     +A_{4}\nabla _{h}^{4}v \label{eq:dissipation}
925     \end{equation}
926     where $A_{h}$ and $A_{v}\ $are (constant) horizontal and vertical viscosity
927     coefficients and $A_{4}\ $is the horizontal coefficient for biharmonic
928     friction. These coefficients are the same for all velocity components.
929    
930     \paragraph{Tracers}
931    
932     The mixing terms for the temperature and salinity equations have a similar
933     form to that of momentum except that the diffusion tensor can be
934     non-diagonal and have varying coefficients. $\qquad $%
935     \begin{equation}
936     D_{T,S}=\nabla .[\underline{\underline{K}}\nabla (T,S)]+K_{4}\nabla
937     _{h}^{4}(T,S) \label{eq:diffusion}
938     \end{equation}
939     where $\underline{\underline{K}}\ $is the diffusion tensor and the $K_{4}\ $%
940     horizontal coefficient for biharmonic diffusion. In the simplest case where
941     the subgrid-scale fluxes of heat and salt are parameterized with constant
942     horizontal and vertical diffusion coefficients, $\underline{\underline{K}}$,
943     reduces to a diagonal matrix with constant coefficients:
944    
945     \begin{equation}
946     \qquad \qquad \qquad \qquad K=\left(
947     \begin{array}{ccc}
948     K_{h} & 0 & 0 \\
949     0 & K_{h} & 0 \\
950     0 & 0 & K_{v}%
951     \end{array}
952     \right) \qquad \qquad \qquad \label{eq:diagonal-diffusion-tensor}
953     \end{equation}
954     where $K_{h}\ $and $K_{v}\ $are the horizontal and vertical diffusion
955     coefficients. These coefficients are the same for all tracers (temperature,
956     salinity ... ).
957    
958     \subsection{Vector invariant form}
959    
960     For some purposes it is advantageous to write momentum advection in eq(\ref%
961     {hor-mtm}) and (\ref{vertmtm}) in the (so-called) `vector invariant' form:
962    
963     \begin{equation}
964     \frac{D\vec{\mathbf{v}}}{Dt}=\frac{\partial \vec{\mathbf{v}}}{\partial t}%
965     +\left( \nabla \times \vec{\mathbf{v}}\right) \times \vec{\mathbf{v}}+\nabla %
966 cnh 1.2 \left[ \frac{1}{2}(\vec{\mathbf{v}}\cdot \vec{\mathbf{v}})\right]
967 cnh 1.1 \label{eq:vi-identity}
968     \end{equation}%
969     This permits alternative numerical treatments of the non-linear terms based
970     on their representation as a vorticity flux. Because gradients of coordinate
971     vectors no longer appear on the rhs of (\ref{eq:vi-identity}), explicit
972     representation of the metric terms in (\ref{eq:gu-speherical}), (\ref%
973     {eq:gv-spherical}) and (\ref{eq:gw-spherical}), can be avoided: information
974     about the geometry is contained in the areas and lengths of the volumes used
975     to discretize the model.
976    
977     \subsection{Adjoint}
978    
979 cnh 1.2 Tangent linear and adjoint counterparts of the forward model and described
980     in Chapter 5.
981 cnh 1.1
982     % $Header: /u/gcmpack/mitgcmdoc/part1/appendix_atmos.tex,v 1.3 2001/09/26 14:53:10 cnh Exp $
983     % $Name: $
984    
985     \section{Appendix ATMOSPHERE}
986    
987     \subsection{Hydrostatic Primitive Equations for the Atmosphere in pressure
988     coordinates}
989    
990     \label{sect-hpe-p}
991    
992     The hydrostatic primitive equations (HPEs) in p-coordinates are:
993     \begin{eqnarray}
994     \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}%
995 cnh 1.2 _{h}+\mathbf{\nabla }_{p}\phi &=&\vec{\mathbf{\mathcal{F}}}
996 cnh 1.1 \label{eq:atmos-mom} \\
997 cnh 1.2 \frac{\partial \phi }{\partial p}+\alpha &=&0 \label{eq-p-hydro-start} \\
998 cnh 1.1 \mathbf{\nabla }_{p}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \omega }{%
999     \partial p} &=&0 \label{eq:atmos-cont} \\
1000 cnh 1.2 p\alpha &=&RT \label{eq:atmos-eos} \\
1001 cnh 1.1 c_{v}\frac{DT}{Dt}+p\frac{D\alpha }{Dt} &=&\mathcal{Q} \label{eq:atmos-heat}
1002     \end{eqnarray}%
1003     where $\vec{\mathbf{v}}_{h}=(u,v,0)$ is the `horizontal' (on pressure
1004     surfaces) component of velocity,$\frac{D}{Dt}=\vec{\mathbf{v}}_{h}\cdot
1005     \mathbf{\nabla }_{p}+\omega \frac{\partial }{\partial p}$ is the total
1006     derivative, $f=2\Omega \sin lat$ is the Coriolis parameter, $\phi =gz$ is
1007     the geopotential, $\alpha =1/\rho $ is the specific volume, $\omega =\frac{Dp%
1008     }{Dt}$ is the vertical velocity in the $p-$coordinate. Equation(\ref%
1009     {eq:atmos-heat}) is the first law of thermodynamics where internal energy $%
1010     e=c_{v}T$, $T$ is temperature, $Q$ is the rate of heating per unit mass and $%
1011     p\frac{D\alpha }{Dt}$ is the work done by the fluid in compressing.
1012    
1013     It is convenient to cast the heat equation in terms of potential temperature
1014     $\theta $ so that it looks more like a generic conservation law.
1015     Differentiating (\ref{eq:atmos-eos}) we get:
1016     \begin{equation*}
1017     p\frac{D\alpha }{Dt}+\alpha \frac{Dp}{Dt}=R\frac{DT}{Dt}
1018     \end{equation*}%
1019     which, when added to the heat equation (\ref{eq:atmos-heat}) and using $%
1020     c_{p}=c_{v}+R$, gives:
1021     \begin{equation}
1022     c_{p}\frac{DT}{Dt}-\alpha \frac{Dp}{Dt}=\mathcal{Q}
1023     \label{eq-p-heat-interim}
1024     \end{equation}%
1025     Potential temperature is defined:
1026     \begin{equation}
1027     \theta =T(\frac{p_{c}}{p})^{\kappa } \label{eq:potential-temp}
1028     \end{equation}%
1029     where $p_{c}$ is a reference pressure and $\kappa =R/c_{p}$. For convenience
1030     we will make use of the Exner function $\Pi (p)$ which defined by:
1031     \begin{equation}
1032     \Pi (p)=c_{p}(\frac{p}{p_{c}})^{\kappa } \label{Exner}
1033     \end{equation}%
1034     The following relations will be useful and are easily expressed in terms of
1035     the Exner function:
1036     \begin{equation*}
1037     c_{p}T=\Pi \theta \;\;;\;\;\frac{\partial \Pi }{\partial p}=\frac{\kappa \Pi
1038     }{p}\;\;;\;\;\alpha =\frac{\kappa \Pi \theta }{p}=\frac{\partial \ \Pi }{%
1039     \partial p}\theta \;\;;\;\;\frac{D\Pi }{Dt}=\frac{\partial \Pi }{\partial p}%
1040     \frac{Dp}{Dt}
1041     \end{equation*}%
1042     where $b=\frac{\partial \ \Pi }{\partial p}\theta $ is the buoyancy.
1043    
1044     The heat equation is obtained by noting that
1045     \begin{equation*}
1046     c_{p}\frac{DT}{Dt}=\frac{D(\Pi \theta )}{Dt}=\Pi \frac{D\theta }{Dt}+\theta
1047 cnh 1.2 \frac{D\Pi }{Dt}=\Pi \frac{D\theta }{Dt}+\alpha \frac{Dp}{Dt}
1048 cnh 1.1 \end{equation*}
1049     and on substituting into (\ref{eq-p-heat-interim}) gives:
1050     \begin{equation}
1051     \Pi \frac{D\theta }{Dt}=\mathcal{Q}
1052     \label{eq:potential-temperature-equation}
1053     \end{equation}
1054     which is in conservative form.
1055    
1056     For convenience in the model we prefer to step forward (\ref%
1057     {eq:potential-temperature-equation}) rather than (\ref{eq:atmos-heat}).
1058    
1059     \subsubsection{Boundary conditions}
1060    
1061     The upper and lower boundary conditions are :
1062     \begin{eqnarray}
1063     \mbox{at the top:}\;\;p=0 &&\text{, }\omega =\frac{Dp}{Dt}=0 \\
1064     \mbox{at the surface:}\;\;p=p_{s} &&\text{, }\phi =\phi _{topo}=g~Z_{topo}
1065     \label{eq:boundary-condition-atmosphere}
1066     \end{eqnarray}
1067     In $p$-coordinates, the upper boundary acts like a solid boundary ($\omega
1068     =0 $); in $z$-coordinates and the lower boundary is analogous to a free
1069     surface ($\phi $ is imposed and $\omega \neq 0$).
1070    
1071     \subsubsection{Splitting the geo-potential}
1072    
1073     For the purposes of initialization and reducing round-off errors, the model
1074     deals with perturbations from reference (or ``standard'') profiles. For
1075     example, the hydrostatic geopotential associated with the resting atmosphere
1076     is not dynamically relevant and can therefore be subtracted from the
1077     equations. The equations written in terms of perturbations are obtained by
1078     substituting the following definitions into the previous model equations:
1079     \begin{eqnarray}
1080     \theta &=&\theta _{o}+\theta ^{\prime } \label{eq:atmos-ref-prof-theta} \\
1081     \alpha &=&\alpha _{o}+\alpha ^{\prime } \label{eq:atmos-ref-prof-alpha} \\
1082     \phi &=&\phi _{o}+\phi ^{\prime } \label{eq:atmos-ref-prof-phi}
1083     \end{eqnarray}
1084     The reference state (indicated by subscript ``0'') corresponds to
1085     horizontally homogeneous atmosphere at rest ($\theta _{o},\alpha _{o},\phi
1086     _{o}$) with surface pressure $p_{o}(x,y)$ that satisfies $\phi
1087     _{o}(p_{o})=g~Z_{topo}$, defined:
1088     \begin{eqnarray*}
1089     \theta _{o}(p) &=&f^{n}(p) \\
1090     \alpha _{o}(p) &=&\Pi _{p}\theta _{o} \\
1091     \phi _{o}(p) &=&\phi _{topo}-\int_{p_{0}}^{p}\alpha _{o}dp
1092     \end{eqnarray*}
1093     %\begin{eqnarray*}
1094     %\phi'_\alpha & = & \int^p_{p_o} (\alpha_o -\alpha) dp \\
1095     %\phi'_s(x,y,t) & = & \int_{p_o}^{p_s} \alpha dp
1096     %\end{eqnarray*}
1097    
1098     The final form of the HPE's in p coordinates is then:
1099     \begin{eqnarray}
1100     \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}%
1101     _{h}+\mathbf{\nabla }_{p}\phi ^{\prime } &=&\vec{\mathbf{\mathcal{F}}} \\
1102     \frac{\partial \phi ^{\prime }}{\partial p}+\alpha ^{\prime } &=&0 \\
1103     \mathbf{\nabla }_{p}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \omega }{%
1104     \partial p} &=&0 \\
1105     \frac{\partial \Pi }{\partial p}\theta ^{\prime } &=&\alpha ^{\prime } \\
1106     \frac{D\theta }{Dt} &=&\frac{\mathcal{Q}}{\Pi } \label{eq:atmos-prime}
1107     \end{eqnarray}
1108    
1109     % $Header: /u/gcmpack/mitgcmdoc/part1/appendix_ocean.tex,v 1.2 2001/09/11 14:39:38 cnh Exp $
1110     % $Name: $
1111    
1112     \section{Appendix OCEAN}
1113    
1114     \subsection{Equations of motion for the ocean}
1115    
1116     We review here the method by which the standard (Boussinesq, incompressible)
1117     HPE's for the ocean written in z-coordinates are obtained. The
1118     non-Boussinesq equations for oceanic motion are:
1119     \begin{eqnarray}
1120     \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}%
1121     _{h}+\frac{1}{\rho }\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}} \\
1122     \epsilon _{nh}\frac{Dw}{Dt}+g+\frac{1}{\rho }\frac{\partial p}{\partial z}
1123     &=&\epsilon _{nh}\mathcal{F}_{w} \\
1124     \frac{1}{\rho }\frac{D\rho }{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}%
1125     _{h}+\frac{\partial w}{\partial z} &=&0 \\
1126 cnh 1.2 \rho &=&\rho (\theta ,S,p) \\
1127 cnh 1.1 \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \\
1128     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq:non-boussinesq}
1129     \end{eqnarray}%
1130     These equations permit acoustics modes, inertia-gravity waves,
1131     non-hydrostatic motions, a geostrophic (Rossby) mode and a thermo-haline
1132     mode. As written, they cannot be integrated forward consistently - if we
1133     step $\rho $ forward in (\ref{eq-zns-cont}), the answer will not be
1134     consistent with that obtained by stepping (\ref{eq-zns-heat}) and (\ref%
1135     {eq-zns-salt}) and then using (\ref{eq-zns-eos}) to yield $\rho $. It is
1136     therefore necessary to manipulate the system as follows. Differentiating the
1137     EOS (equation of state) gives:
1138    
1139     \begin{equation}
1140     \frac{D\rho }{Dt}=\left. \frac{\partial \rho }{\partial \theta }\right|
1141     _{S,p}\frac{D\theta }{Dt}+\left. \frac{\partial \rho }{\partial S}\right|
1142     _{\theta ,p}\frac{DS}{Dt}+\left. \frac{\partial \rho }{\partial p}\right|
1143     _{\theta ,S}\frac{Dp}{Dt} \label{EOSexpansion}
1144     \end{equation}
1145    
1146     Note that $\frac{\partial \rho }{\partial p}=\frac{1}{c_{s}^{2}}$ is the
1147     reciprocal of the sound speed ($c_{s}$) squared. Substituting into \ref%
1148     {eq-zns-cont} gives:
1149     \begin{equation}
1150     \frac{1}{\rho c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{%
1151     v}}+\partial _{z}w\approx 0 \label{eq-zns-pressure}
1152     \end{equation}
1153     where we have used an approximation sign to indicate that we have assumed
1154     adiabatic motion, dropping the $\frac{D\theta }{Dt}$ and $\frac{DS}{Dt}$.
1155     Replacing \ref{eq-zns-cont} with \ref{eq-zns-pressure} yields a system that
1156     can be explicitly integrated forward:
1157     \begin{eqnarray}
1158     \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}%
1159     _{h}+\frac{1}{\rho }\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1160     \label{eq-cns-hmom} \\
1161     \epsilon _{nh}\frac{Dw}{Dt}+g+\frac{1}{\rho }\frac{\partial p}{\partial z}
1162     &=&\epsilon _{nh}\mathcal{F}_{w} \label{eq-cns-hydro} \\
1163     \frac{1}{\rho c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{%
1164     v}}_{h}+\frac{\partial w}{\partial z} &=&0 \label{eq-cns-cont} \\
1165     \rho &=&\rho (\theta ,S,p) \label{eq-cns-eos} \\
1166     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-cns-heat} \\
1167     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-cns-salt}
1168     \end{eqnarray}
1169    
1170     \subsubsection{Compressible z-coordinate equations}
1171    
1172     Here we linearize the acoustic modes by replacing $\rho $ with $\rho _{o}(z)$
1173     wherever it appears in a product (ie. non-linear term) - this is the
1174     `Boussinesq assumption'. The only term that then retains the full variation
1175     in $\rho $ is the gravitational acceleration:
1176     \begin{eqnarray}
1177     \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}%
1178     _{h}+\frac{1}{\rho _{o}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1179     \label{eq-zcb-hmom} \\
1180     \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{o}}+\frac{1}{\rho _{o}}%
1181     \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1182     \label{eq-zcb-hydro} \\
1183     \frac{1}{\rho _{o}c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{%
1184     \mathbf{v}}_{h}+\frac{\partial w}{\partial z} &=&0 \label{eq-zcb-cont} \\
1185     \rho &=&\rho (\theta ,S,p) \label{eq-zcb-eos} \\
1186     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-zcb-heat} \\
1187     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-zcb-salt}
1188     \end{eqnarray}
1189     These equations still retain acoustic modes. But, because the
1190     ``compressible'' terms are linearized, the pressure equation \ref%
1191     {eq-zcb-cont} can be integrated implicitly with ease (the time-dependent
1192     term appears as a Helmholtz term in the non-hydrostatic pressure equation).
1193     These are the \emph{truly} compressible Boussinesq equations. Note that the
1194     EOS must have the same pressure dependency as the linearized pressure term,
1195     ie. $\left. \frac{\partial \rho }{\partial p}\right| _{\theta ,S}=\frac{1}{%
1196     c_{s}^{2}}$, for consistency.
1197    
1198     \subsubsection{`Anelastic' z-coordinate equations}
1199    
1200     The anelastic approximation filters the acoustic mode by removing the
1201     time-dependency in the continuity (now pressure-) equation (\ref{eq-zcb-cont}%
1202     ). This could be done simply by noting that $\frac{Dp}{Dt}\approx -g\rho _{o}%
1203     \frac{Dz}{Dt}=-g\rho _{o}w$, but this leads to an inconsistency between
1204     continuity and EOS. A better solution is to change the dependency on
1205     pressure in the EOS by splitting the pressure into a reference function of
1206     height and a perturbation:
1207     \begin{equation*}
1208 cnh 1.2 \rho =\rho (\theta ,S,p_{o}(z)+\epsilon _{s}p^{\prime })
1209 cnh 1.1 \end{equation*}
1210     Remembering that the term $\frac{Dp}{Dt}$ in continuity comes from
1211     differentiating the EOS, the continuity equation then becomes:
1212     \begin{equation*}
1213     \frac{1}{\rho _{o}c_{s}^{2}}\left( \frac{Dp_{o}}{Dt}+\epsilon _{s}\frac{%
1214     Dp^{\prime }}{Dt}\right) +\mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+%
1215 cnh 1.2 \frac{\partial w}{\partial z}=0
1216 cnh 1.1 \end{equation*}
1217     If the time- and space-scales of the motions of interest are longer than
1218     those of acoustic modes, then $\frac{Dp^{\prime }}{Dt}<<(\frac{Dp_{o}}{Dt},%
1219     \mathbf{\nabla }\cdot \vec{\mathbf{v}}_{h})$ in the continuity equations and
1220     $\left. \frac{\partial \rho }{\partial p}\right| _{\theta ,S}\frac{%
1221     Dp^{\prime }}{Dt}<<\left. \frac{\partial \rho }{\partial p}\right| _{\theta
1222     ,S}\frac{Dp_{o}}{Dt}$ in the EOS (\ref{EOSexpansion}). Thus we set $\epsilon
1223     _{s}=0$, removing the dependency on $p^{\prime }$ in the continuity equation
1224     and EOS. Expanding $\frac{Dp_{o}(z)}{Dt}=-g\rho _{o}w$ then leads to the
1225     anelastic continuity equation:
1226     \begin{equation}
1227     \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z}-%
1228     \frac{g}{c_{s}^{2}}w=0 \label{eq-za-cont1}
1229     \end{equation}
1230     A slightly different route leads to the quasi-Boussinesq continuity equation
1231     where we use the scaling $\frac{\partial \rho ^{\prime }}{\partial t}+%
1232     \mathbf{\nabla }_{3}\cdot \rho ^{\prime }\vec{\mathbf{v}}<<\mathbf{\nabla }%
1233     _{3}\cdot \rho _{o}\vec{\mathbf{v}}$ yielding:
1234     \begin{equation}
1235     \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{1}{\rho _{o}}\frac{%
1236     \partial \left( \rho _{o}w\right) }{\partial z}=0 \label{eq-za-cont2}
1237     \end{equation}
1238     Equations \ref{eq-za-cont1} and \ref{eq-za-cont2} are in fact the same
1239     equation if:
1240     \begin{equation}
1241     \frac{1}{\rho _{o}}\frac{\partial \rho _{o}}{\partial z}=\frac{-g}{c_{s}^{2}}
1242     \end{equation}
1243     Again, note that if $\rho _{o}$ is evaluated from prescribed $\theta _{o}$
1244     and $S_{o}$ profiles, then the EOS dependency on $p_{o}$ and the term $\frac{%
1245     g}{c_{s}^{2}}$ in continuity should be referred to those same profiles. The
1246     full set of `quasi-Boussinesq' or `anelastic' equations for the ocean are
1247     then:
1248     \begin{eqnarray}
1249     \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}%
1250     _{h}+\frac{1}{\rho _{o}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1251     \label{eq-zab-hmom} \\
1252     \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{o}}+\frac{1}{\rho _{o}}%
1253     \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1254     \label{eq-zab-hydro} \\
1255     \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{1}{\rho _{o}}\frac{%
1256     \partial \left( \rho _{o}w\right) }{\partial z} &=&0 \label{eq-zab-cont} \\
1257     \rho &=&\rho (\theta ,S,p_{o}(z)) \label{eq-zab-eos} \\
1258     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-zab-heat} \\
1259     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-zab-salt}
1260     \end{eqnarray}
1261    
1262     \subsubsection{Incompressible z-coordinate equations}
1263    
1264     Here, the objective is to drop the depth dependence of $\rho _{o}$ and so,
1265     technically, to also remove the dependence of $\rho $ on $p_{o}$. This would
1266     yield the ``truly'' incompressible Boussinesq equations:
1267     \begin{eqnarray}
1268     \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}%
1269     _{h}+\frac{1}{\rho _{c}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1270     \label{eq-ztb-hmom} \\
1271     \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{c}}+\frac{1}{\rho _{c}}%
1272     \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1273     \label{eq-ztb-hydro} \\
1274     \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z}
1275     &=&0 \label{eq-ztb-cont} \\
1276     \rho &=&\rho (\theta ,S) \label{eq-ztb-eos} \\
1277     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-ztb-heat} \\
1278     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-ztb-salt}
1279     \end{eqnarray}
1280     where $\rho _{c}$ is a constant reference density of water.
1281    
1282     \subsubsection{Compressible non-divergent equations}
1283    
1284     The above ``incompressible'' equations are incompressible in both the flow
1285     and the density. In many oceanic applications, however, it is important to
1286     retain compressibility effects in the density. To do this we must split the
1287     density thus:
1288     \begin{equation*}
1289     \rho =\rho _{o}+\rho ^{\prime }
1290     \end{equation*}%
1291     We then assert that variations with depth of $\rho _{o}$ are unimportant
1292     while the compressible effects in $\rho ^{\prime }$ are:
1293     \begin{equation*}
1294     \rho _{o}=\rho _{c}
1295     \end{equation*}%
1296     \begin{equation*}
1297     \rho ^{\prime }=\rho (\theta ,S,p_{o}(z))-\rho _{o}
1298     \end{equation*}%
1299     This then yields what we can call the semi-compressible Boussinesq
1300     equations:
1301     \begin{eqnarray}
1302     \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}%
1303     _{h}+\frac{1}{\rho _{c}}\mathbf{\nabla }_{z}p^{\prime } &=&\vec{\mathbf{%
1304     \mathcal{F}}} \label{eq:ocean-mom} \\
1305     \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho ^{\prime }}{\rho _{c}}+\frac{1}{\rho
1306     _{c}}\frac{\partial p^{\prime }}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1307     \label{eq:ocean-wmom} \\
1308     \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z}
1309     &=&0 \label{eq:ocean-cont} \\
1310     \rho ^{\prime } &=&\rho (\theta ,S,p_{o}(z))-\rho _{c} \label{eq:ocean-eos}
1311     \\
1312     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq:ocean-theta} \\
1313     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq:ocean-salt}
1314     \end{eqnarray}%
1315     Note that the hydrostatic pressure of the resting fluid, including that
1316     associated with $\rho _{c}$, is subtracted out since it has no effect on the
1317     dynamics.
1318    
1319     Though necessary, the assumptions that go into these equations are messy
1320     since we essentially assume a different EOS for the reference density and
1321     the perturbation density. Nevertheless, it is the hydrostatic ($\epsilon
1322     _{nh}=0$ form of these equations that are used throughout the ocean modeling
1323     community and referred to as the primitive equations (HPE).
1324    
1325     % $Header: /u/gcmpack/mitgcmdoc/part1/appendix_operators.tex,v 1.1.1.1 2001/08/08 16:16:19 adcroft Exp $
1326     % $Name: $
1327    
1328     \section{Appendix:OPERATORS}
1329    
1330     \subsection{Coordinate systems}
1331    
1332     \subsubsection{Spherical coordinates}
1333    
1334     In spherical coordinates, the velocity components in the zonal, meridional
1335     and vertical direction respectively, are given by (see Fig.2) :
1336    
1337     \begin{equation*}
1338 cnh 1.2 u=r\cos \phi \frac{D\lambda }{Dt}
1339 cnh 1.1 \end{equation*}
1340    
1341     \begin{equation*}
1342 cnh 1.2 v=r\frac{D\phi }{Dt}\qquad
1343 cnh 1.1 \end{equation*}
1344     $\qquad \qquad \qquad \qquad $
1345    
1346     \begin{equation*}
1347 cnh 1.2 \dot{r}=\frac{Dr}{Dt}
1348 cnh 1.1 \end{equation*}
1349    
1350     Here $\phi $ is the latitude, $\lambda $ the longitude, $r$ the radial
1351     distance of the particle from the center of the earth, $\Omega $ is the
1352     angular speed of rotation of the Earth and $D/Dt$ is the total derivative.
1353    
1354     The `grad' ($\nabla $) and `div' ($\nabla $.) operators are defined by, in
1355     spherical coordinates:
1356    
1357     \begin{equation*}
1358     \nabla \equiv \left( \frac{1}{r\cos \phi }\frac{\partial }{\partial \lambda }%
1359     ,\frac{1}{r}\frac{\partial }{\partial \phi },\frac{\partial }{\partial r}%
1360 cnh 1.2 \right)
1361 cnh 1.1 \end{equation*}
1362    
1363     \begin{equation*}
1364     \nabla .v\equiv \frac{1}{r\cos \phi }\left\{ \frac{\partial u}{\partial
1365     \lambda }+\frac{\partial }{\partial \phi }\left( v\cos \phi \right) \right\}
1366 cnh 1.2 +\frac{1}{r^{2}}\frac{\partial \left( r^{2}\dot{r}\right) }{\partial r}
1367 cnh 1.1 \end{equation*}
1368    
1369     %%%% \end{document}

  ViewVC Help
Powered by ViewVC 1.1.22