/[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 1.7 - (hide annotations) (download) (as text)
Thu Oct 25 12:06:56 2001 UTC (23 years, 8 months ago) by cnh
Branch: MAIN
Changes since 1.6: +65 -58 lines
File MIME type: application/x-tex
More editing

1 cnh 1.7 % $Header: /u/u0/gcmpack/mitgcmdoc/part1/manual.tex,v 1.6 2001/10/24 15:21:27 cnh Exp $
2 cnh 1.2 % $Name: $
3 cnh 1.1
4 adcroft 1.4 %tci%\documentclass[12pt]{book}
5     %tci%\usepackage{amsmath}
6     %tci%\usepackage{html}
7     %tci%\usepackage{epsfig}
8     %tci%\usepackage{graphics,subfigure}
9     %tci%\usepackage{array}
10     %tci%\usepackage{multirow}
11     %tci%\usepackage{fancyhdr}
12     %tci%\usepackage{psfrag}
13    
14     %tci%%TCIDATA{OutputFilter=Latex.dll}
15     %tci%%TCIDATA{LastRevised=Thursday, October 04, 2001 14:41:22}
16     %tci%%TCIDATA{<META NAME="GraphicsSave" CONTENT="32">}
17     %tci%%TCIDATA{Language=American English}
18    
19     %tci%\fancyhead{}
20     %tci%\fancyhead[LO]{\slshape \rightmark}
21     %tci%\fancyhead[RE]{\slshape \leftmark}
22     %tci%\fancyhead[RO,LE]{\thepage}
23     %tci%\fancyfoot[CO,CE]{\today}
24     %tci%\fancyfoot[RO,LE]{ }
25     %tci%\renewcommand{\headrulewidth}{0.4pt}
26     %tci%\renewcommand{\footrulewidth}{0.4pt}
27     %tci%\setcounter{secnumdepth}{3}
28     %tci%\input{tcilatex}
29    
30     %tci%\begin{document}
31    
32     %tci%\tableofcontents
33    
34    
35 cnh 1.1 % Section: Overview
36    
37 cnh 1.7 % $Header: /u/u0/gcmpack/mitgcmdoc/part1/manual.tex,v 1.6 2001/10/24 15:21:27 cnh Exp $
38 cnh 1.1 % $Name: $
39    
40     \section{Introduction}
41    
42     This documentation provides the reader with the information necessary to
43     carry out numerical experiments using MITgcm. It gives a comprehensive
44     description of the continuous equations on which the model is based, the
45     numerical algorithms the model employs and a description of the associated
46     program code. Along with the hydrodynamical kernel, physical and
47     biogeochemical parameterizations of key atmospheric and oceanic processes
48     are available. A number of examples illustrating the use of the model in
49     both process and general circulation studies of the atmosphere and ocean are
50     also presented.
51    
52     MITgcm has a number of novel aspects:
53    
54     \begin{itemize}
55     \item it can be used to study both atmospheric and oceanic phenomena; one
56     hydrodynamical kernel is used to drive forward both atmospheric and oceanic
57 cnh 1.7 models - see fig \ref{fig:onemodel}
58 cnh 1.1
59 cnh 1.3 %% CNHbegin
60     \input{part1/one_model_figure}
61     %% CNHend
62    
63 cnh 1.1 \item it has a non-hydrostatic capability and so can be used to study both
64 cnh 1.7 small-scale and large scale processes - see fig \ref{fig:all-scales}
65 cnh 1.1
66 cnh 1.3 %% CNHbegin
67     \input{part1/all_scales_figure}
68     %% CNHend
69    
70 cnh 1.1 \item finite volume techniques are employed yielding an intuitive
71     discretization and support for the treatment of irregular geometries using
72 cnh 1.7 orthogonal curvilinear grids and shaved cells - see fig \ref{fig:finite-volumes}
73 cnh 1.3
74     %% CNHbegin
75     \input{part1/fvol_figure}
76     %% CNHend
77 cnh 1.1
78     \item tangent linear and adjoint counterparts are automatically maintained
79     along with the forward model, permitting sensitivity and optimization
80     studies.
81    
82     \item the model is developed to perform efficiently on a wide variety of
83     computational platforms.
84     \end{itemize}
85    
86     Key publications reporting on and charting the development of the model are
87     listed in an Appendix.
88    
89     We begin by briefly showing some of the results of the model in action to
90     give a feel for the wide range of problems that can be addressed using it.
91    
92 cnh 1.7 % $Header: /u/u0/gcmpack/mitgcmdoc/part1/manual.tex,v 1.6 2001/10/24 15:21:27 cnh Exp $
93 cnh 1.1 % $Name: $
94    
95     \section{Illustrations of the model in action}
96    
97     The MITgcm has been designed and used to model a wide range of phenomena,
98     from convection on the scale of meters in the ocean to the global pattern of
99 cnh 1.7 atmospheric winds - see figure \ref{fig:all-scales}. To give a flavor of the
100 cnh 1.1 kinds of problems the model has been used to study, we briefly describe some
101     of them here. A more detailed description of the underlying formulation,
102     numerical algorithm and implementation that lie behind these calculations is
103 cnh 1.2 given later. Indeed many of the illustrative examples shown below can be
104     easily reproduced: simply download the model (the minimum you need is a PC
105     running linux, together with a FORTRAN\ 77 compiler) and follow the examples
106     described in detail in the documentation.
107 cnh 1.1
108     \subsection{Global atmosphere: `Held-Suarez' benchmark}
109    
110 cnh 1.7 A novel feature of MITgcm is its ability to simulate, using one basic algorithm,
111     both atmospheric and oceanographic flows at both small and large scales.
112 cnh 1.2
113 cnh 1.7 Figure \ref{fig:eddy_cs} shows an instantaneous plot of the 500$mb$
114 cnh 1.2 temperature field obtained using the atmospheric isomorph of MITgcm run at
115     2.8$^{\circ }$ resolution on the cubed sphere. We see cold air over the pole
116     (blue) and warm air along an equatorial band (red). Fully developed
117     baroclinic eddies spawned in the northern hemisphere storm track are
118     evident. There are no mountains or land-sea contrast in this calculation,
119     but you can easily put them in. The model is driven by relaxation to a
120     radiative-convective equilibrium profile, following the description set out
121     in Held and Suarez; 1994 designed to test atmospheric hydrodynamical cores -
122     there are no mountains or land-sea contrast.
123    
124 cnh 1.3 %% CNHbegin
125     \input{part1/cubic_eddies_figure}
126     %% CNHend
127    
128 cnh 1.2 As described in Adcroft (2001), a `cubed sphere' is used to discretize the
129 cnh 1.7 globe permitting a uniform gridding and obviated the need to Fourier filter.
130 cnh 1.2 The `vector-invariant' form of MITgcm supports any orthogonal curvilinear
131     grid, of which the cubed sphere is just one of many choices.
132 cnh 1.1
133 cnh 1.7 Figure \ref{fig:hs_zave_u} shows the 5-year mean, zonally averaged zonal
134     wind from a 20-level configuration of
135 cnh 1.2 the model. It compares favorable with more conventional spatial
136 cnh 1.7 discretization approaches. The two plots show the field calculated using the
137     cube-sphere grid and the flow calculated using a regular, spherical polar
138     latitude-longitude grid. Both grids are supported within the model.
139 cnh 1.1
140 cnh 1.3 %% CNHbegin
141     \input{part1/hs_zave_u_figure}
142     %% CNHend
143    
144 cnh 1.2 \subsection{Ocean gyres}
145 cnh 1.1
146 cnh 1.2 Baroclinic instability is a ubiquitous process in the ocean, as well as the
147     atmosphere. Ocean eddies play an important role in modifying the
148     hydrographic structure and current systems of the oceans. Coarse resolution
149     models of the oceans cannot resolve the eddy field and yield rather broad,
150     diffusive patterns of ocean currents. But if the resolution of our models is
151     increased until the baroclinic instability process is resolved, numerical
152     solutions of a different and much more realistic kind, can be obtained.
153    
154 cnh 1.7 Figure \ref{fig:ocean-gyres} shows the surface temperature and velocity
155     field obtained from MITgcm run at $\frac{1}{6}^{\circ }$ horizontal
156     resolution on a $lat-lon$
157 cnh 1.2 grid in which the pole has been rotated by 90$^{\circ }$ on to the equator
158     (to avoid the converging of meridian in northern latitudes). 21 vertical
159     levels are used in the vertical with a `lopped cell' representation of
160     topography. The development and propagation of anomalously warm and cold
161 cnh 1.7 eddies can be clearly seen in the Gulf Stream region. The transport of
162 cnh 1.2 warm water northward by the mean flow of the Gulf Stream is also clearly
163     visible.
164 cnh 1.1
165 cnh 1.3 %% CNHbegin
166     \input{part1/ocean_gyres_figure}
167     %% CNHend
168    
169    
170 cnh 1.1 \subsection{Global ocean circulation}
171    
172 cnh 1.7 Figure \ref{fig:large-scale-circ} (top) shows the pattern of ocean currents at
173     the surface of a 4$^{\circ }$
174 cnh 1.2 global ocean model run with 15 vertical levels. Lopped cells are used to
175     represent topography on a regular $lat-lon$ grid extending from 70$^{\circ
176     }N $ to 70$^{\circ }S$. The model is driven using monthly-mean winds with
177     mixed boundary conditions on temperature and salinity at the surface. The
178     transfer properties of ocean eddies, convection and mixing is parameterized
179     in this model.
180    
181 cnh 1.7 Figure \ref{fig:large-scale-circ} (bottom) shows the meridional overturning
182     circulation of the global ocean in Sverdrups.
183 cnh 1.2
184 cnh 1.3 %%CNHbegin
185     \input{part1/global_circ_figure}
186     %%CNHend
187    
188 cnh 1.2 \subsection{Convection and mixing over topography}
189    
190     Dense plumes generated by localized cooling on the continental shelf of the
191     ocean may be influenced by rotation when the deformation radius is smaller
192     than the width of the cooling region. Rather than gravity plumes, the
193     mechanism for moving dense fluid down the shelf is then through geostrophic
194 cnh 1.7 eddies. The simulation shown in the figure \ref{fig::convect-and-topo}
195     (blue is cold dense fluid, red is
196 cnh 1.2 warmer, lighter fluid) employs the non-hydrostatic capability of MITgcm to
197     trigger convection by surface cooling. The cold, dense water falls down the
198     slope but is deflected along the slope by rotation. It is found that
199     entrainment in the vertical plane is reduced when rotational control is
200     strong, and replaced by lateral entrainment due to the baroclinic
201     instability of the along-slope current.
202 cnh 1.1
203 cnh 1.3 %%CNHbegin
204     \input{part1/convect_and_topo}
205     %%CNHend
206    
207 cnh 1.1 \subsection{Boundary forced internal waves}
208    
209 cnh 1.2 The unique ability of MITgcm to treat non-hydrostatic dynamics in the
210     presence of complex geometry makes it an ideal tool to study internal wave
211     dynamics and mixing in oceanic canyons and ridges driven by large amplitude
212     barotropic tidal currents imposed through open boundary conditions.
213    
214 cnh 1.7 Fig. \ref{fig:boundary-forced-wave} shows the influence of cross-slope
215     topographic variations on
216 cnh 1.2 internal wave breaking - the cross-slope velocity is in color, the density
217     contoured. The internal waves are excited by application of open boundary
218 cnh 1.7 conditions on the left. They propagate to the sloping boundary (represented
219 cnh 1.2 using MITgcm's finite volume spatial discretization) where they break under
220     nonhydrostatic dynamics.
221    
222 cnh 1.3 %%CNHbegin
223     \input{part1/boundary_forced_waves}
224     %%CNHend
225    
226 cnh 1.2 \subsection{Parameter sensitivity using the adjoint of MITgcm}
227    
228     Forward and tangent linear counterparts of MITgcm are supported using an
229     `automatic adjoint compiler'. These can be used in parameter sensitivity and
230     data assimilation studies.
231    
232 cnh 1.7 As one example of application of the MITgcm adjoint, Figure \ref{fig:hf-sensitivity}
233     maps the gradient $\frac{\partial J}{\partial \mathcal{H}}$where $J$ is the magnitude
234     of the overturning streamfunction shown in figure \ref{fig:large-scale-circ}
235     at 60$^{\circ }$N and $
236     \mathcal{H}(\lambda,\varphi)$ is the mean, local air-sea heat flux over
237     a 100 year period. We see that $J$ is
238 cnh 1.2 sensitive to heat fluxes over the Labrador Sea, one of the important sources
239     of deep water for the thermohaline circulations. This calculation also
240     yields sensitivities to all other model parameters.
241    
242 cnh 1.3 %%CNHbegin
243     \input{part1/adj_hf_ocean_figure}
244     %%CNHend
245    
246 cnh 1.2 \subsection{Global state estimation of the ocean}
247    
248     An important application of MITgcm is in state estimation of the global
249     ocean circulation. An appropriately defined `cost function', which measures
250     the departure of the model from observations (both remotely sensed and
251     insitu) over an interval of time, is minimized by adjusting `control
252     parameters' such as air-sea fluxes, the wind field, the initial conditions
253 cnh 1.7 etc. Figure \ref{fig:assimilated-globes} shows an estimate of the time-mean
254     surface elevation of the ocean obtained by bringing the model in to
255     consistency with altimetric and in-situ observations over the period
256     1992-1997. {\bf CHANGE THIS TEXT - FIG FROM PATRICK/CARL/DETLEF}
257 cnh 1.2
258 cnh 1.3 %% CNHbegin
259     \input{part1/globes_figure}
260     %% CNHend
261    
262 cnh 1.2 \subsection{Ocean biogeochemical cycles}
263    
264     MITgcm is being used to study global biogeochemical cycles in the ocean. For
265     example one can study the effects of interannual changes in meteorological
266     forcing and upper ocean circulation on the fluxes of carbon dioxide and
267 cnh 1.7 oxygen between the ocean and atmosphere. Figure \ref{fig:biogeo} shows
268     the annual air-sea flux of oxygen and its relation to density outcrops in
269     the southern oceans from a single year of a global, interannually varying
270     simulation. The simulation is run at $1^{\circ}\times1^{\circ}$ resolution
271     telescoping to $\frac{1}{3}^{\circ}\times\frac{1}{3}^{\circ}$ in the tropics (not shown).
272 cnh 1.2
273 cnh 1.3 %%CNHbegin
274     \input{part1/biogeo_figure}
275     %%CNHend
276 cnh 1.2
277     \subsection{Simulations of laboratory experiments}
278    
279 cnh 1.7 Figure \ref{fig:lab-simulation} shows MITgcm being used to simulate a
280     laboratory experiment enquiring in to the dynamics of the Antarctic Circumpolar Current (ACC). An
281 cnh 1.2 initially homogeneous tank of water ($1m$ in diameter) is driven from its
282     free surface by a rotating heated disk. The combined action of mechanical
283     and thermal forcing creates a lens of fluid which becomes baroclinically
284     unstable. The stratification and depth of penetration of the lens is
285 cnh 1.7 arrested by its instability in a process analogous to that which sets the
286 cnh 1.2 stratification of the ACC.
287 cnh 1.1
288 cnh 1.3 %%CNHbegin
289     \input{part1/lab_figure}
290     %%CNHend
291    
292 cnh 1.7 % $Header: /u/u0/gcmpack/mitgcmdoc/part1/manual.tex,v 1.6 2001/10/24 15:21:27 cnh Exp $
293 cnh 1.1 % $Name: $
294    
295     \section{Continuous equations in `r' coordinates}
296    
297     To render atmosphere and ocean models from one dynamical core we exploit
298     `isomorphisms' between equation sets that govern the evolution of the
299 cnh 1.7 respective fluids - see figure \ref{fig:isomorphic-equations}.
300     One system of hydrodynamical equations is written down
301 cnh 1.1 and encoded. The model variables have different interpretations depending on
302     whether the atmosphere or ocean is being studied. Thus, for example, the
303     vertical coordinate `$r$' is interpreted as pressure, $p$, if we are
304 cnh 1.7 modeling the atmosphere (left hand side of figure \ref{fig:isomorphic-equations})
305     and height, $z$, if we are modeling the ocean (right hand side of figure
306     \ref{fig:isomorphic-equations}).
307 cnh 1.1
308 cnh 1.3 %%CNHbegin
309     \input{part1/zandpcoord_figure.tex}
310     %%CNHend
311    
312 cnh 1.1 The state of the fluid at any time is characterized by the distribution of
313     velocity $\vec{\mathbf{v}}$, active tracers $\theta $ and $S$, a
314     `geopotential' $\phi $ and density $\rho =\rho (\theta ,S,p)$ which may
315     depend on $\theta $, $S$, and $p$. The equations that govern the evolution
316     of these fields, obtained by applying the laws of classical mechanics and
317     thermodynamics to a Boussinesq, Navier-Stokes fluid are, written in terms of
318 cnh 1.7 a generic vertical coordinate, $r$, so that the appropriate
319     kinematic boundary conditions can be applied isomorphically
320     see figure \ref{fig:zandp-vert-coord}.
321 cnh 1.1
322 cnh 1.3 %%CNHbegin
323     \input{part1/vertcoord_figure.tex}
324     %%CNHend
325    
326 cnh 1.1 \begin{equation*}
327 adcroft 1.4 \frac{D\vec{\mathbf{v}_{h}}}{Dt}+\left( 2\vec{\Omega}\times \vec{\mathbf{v}}
328     \right) _{h}+\mathbf{\nabla }_{h}\phi =\mathcal{F}_{\vec{\mathbf{v}_{h}}}
329 cnh 1.1 \text{ horizontal mtm}
330     \end{equation*}
331    
332     \begin{equation*}
333 adcroft 1.4 \frac{D\dot{r}}{Dt}+\widehat{k}\cdot \left( 2\vec{\Omega}\times \vec{\mathbf{
334 cnh 1.1 v}}\right) +\frac{\partial \phi }{\partial r}+b=\mathcal{F}_{\dot{r}}\text{
335     vertical mtm}
336     \end{equation*}
337    
338     \begin{equation}
339 adcroft 1.4 \mathbf{\nabla }_{h}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \dot{r}}{
340 cnh 1.1 \partial r}=0\text{ continuity} \label{eq:continuous}
341     \end{equation}
342    
343     \begin{equation*}
344 cnh 1.2 b=b(\theta ,S,r)\text{ equation of state}
345 cnh 1.1 \end{equation*}
346    
347     \begin{equation*}
348 cnh 1.2 \frac{D\theta }{Dt}=\mathcal{Q}_{\theta }\text{ potential temperature}
349 cnh 1.1 \end{equation*}
350    
351     \begin{equation*}
352 cnh 1.2 \frac{DS}{Dt}=\mathcal{Q}_{S}\text{ humidity/salinity}
353 cnh 1.1 \end{equation*}
354    
355     Here:
356    
357     \begin{equation*}
358 cnh 1.2 r\text{ is the vertical coordinate}
359 cnh 1.1 \end{equation*}
360    
361     \begin{equation*}
362     \frac{D}{Dt}=\frac{\partial }{\partial t}+\vec{\mathbf{v}}\cdot \nabla \text{
363 cnh 1.2 is the total derivative}
364 cnh 1.1 \end{equation*}
365    
366     \begin{equation*}
367 adcroft 1.4 \mathbf{\nabla }=\mathbf{\nabla }_{h}+\widehat{k}\frac{\partial }{\partial r}
368 cnh 1.2 \text{ is the `grad' operator}
369 cnh 1.1 \end{equation*}
370 adcroft 1.4 with $\mathbf{\nabla }_{h}$ operating in the horizontal and $\widehat{k}
371 cnh 1.1 \frac{\partial }{\partial r}$ operating in the vertical, where $\widehat{k}$
372     is a unit vector in the vertical
373    
374     \begin{equation*}
375 cnh 1.2 t\text{ is time}
376 cnh 1.1 \end{equation*}
377    
378     \begin{equation*}
379     \vec{\mathbf{v}}=(u,v,\dot{r})=(\vec{\mathbf{v}}_{h},\dot{r})\text{ is the
380 cnh 1.2 velocity}
381 cnh 1.1 \end{equation*}
382    
383     \begin{equation*}
384 cnh 1.2 \phi \text{ is the `pressure'/`geopotential'}
385 cnh 1.1 \end{equation*}
386    
387     \begin{equation*}
388 cnh 1.2 \vec{\Omega}\text{ is the Earth's rotation}
389 cnh 1.1 \end{equation*}
390    
391     \begin{equation*}
392 cnh 1.2 b\text{ is the `buoyancy'}
393 cnh 1.1 \end{equation*}
394    
395     \begin{equation*}
396 cnh 1.2 \theta \text{ is potential temperature}
397 cnh 1.1 \end{equation*}
398    
399     \begin{equation*}
400 cnh 1.2 S\text{ is specific humidity in the atmosphere; salinity in the ocean}
401 cnh 1.1 \end{equation*}
402    
403     \begin{equation*}
404 adcroft 1.4 \mathcal{F}_{\vec{\mathbf{v}}}\text{ are forcing and dissipation of }\vec{
405 cnh 1.1 \mathbf{v}}
406     \end{equation*}
407    
408     \begin{equation*}
409 cnh 1.2 \mathcal{Q}_{\theta }\mathcal{\ }\text{are forcing and dissipation of }\theta
410 cnh 1.1 \end{equation*}
411    
412     \begin{equation*}
413     \mathcal{Q}_{S}\mathcal{\ }\text{are forcing and dissipation of }S
414     \end{equation*}
415    
416     The $\mathcal{F}^{\prime }s$ and $\mathcal{Q}^{\prime }s$ are provided by
417 cnh 1.7 `physics' and forcing packages for atmosphere and ocean. These are described
418     in later chapters.
419 cnh 1.1
420     \subsection{Kinematic Boundary conditions}
421    
422     \subsubsection{vertical}
423    
424 cnh 1.7 at fixed and moving $r$ surfaces we set (see figure \ref{fig:zandp-vert-coord}):
425 cnh 1.1
426     \begin{equation}
427     \dot{r}=0atr=R_{fixed}(x,y)\text{ (ocean bottom, top of the atmosphere)}
428     \label{eq:fixedbc}
429     \end{equation}
430    
431     \begin{equation}
432     \dot{r}=\frac{Dr}{Dt}atr=R_{moving}\text{ \
433     (oceansurface,bottomoftheatmosphere)} \label{eq:movingbc}
434     \end{equation}
435    
436     Here
437    
438     \begin{equation*}
439 cnh 1.2 R_{moving}=R_{o}+\eta
440 cnh 1.1 \end{equation*}
441     where $R_{o}(x,y)$ is the `$r-$value' (height or pressure, depending on
442     whether we are in the atmosphere or ocean) of the `moving surface' in the
443     resting fluid and $\eta $ is the departure from $R_{o}(x,y)$ in the presence
444     of motion.
445    
446     \subsubsection{horizontal}
447    
448     \begin{equation}
449     \vec{\mathbf{v}}\cdot \vec{\mathbf{n}}=0 \label{eq:noflow}
450 adcroft 1.4 \end{equation}
451 cnh 1.1 where $\vec{\mathbf{n}}$ is the normal to a solid boundary.
452    
453     \subsection{Atmosphere}
454    
455 cnh 1.7 In the atmosphere, (see figure \ref{fig:zandp-vert-coord}), we interpret:
456 cnh 1.1
457     \begin{equation}
458     r=p\text{ is the pressure} \label{eq:atmos-r}
459     \end{equation}
460    
461     \begin{equation}
462     \dot{r}=\frac{Dp}{Dt}=\omega \text{ is the vertical velocity in }p\text{
463     coordinates} \label{eq:atmos-omega}
464     \end{equation}
465    
466     \begin{equation}
467     \phi =g\,z\text{ is the geopotential height} \label{eq:atmos-phi}
468     \end{equation}
469    
470     \begin{equation}
471     b=\frac{\partial \Pi }{\partial p}\theta \text{ is the buoyancy}
472     \label{eq:atmos-b}
473     \end{equation}
474    
475     \begin{equation}
476     \theta =T(\frac{p_{c}}{p})^{\kappa }\text{ is potential temperature}
477     \label{eq:atmos-theta}
478     \end{equation}
479    
480     \begin{equation}
481     S=q,\text{ is the specific humidity} \label{eq:atmos-s}
482     \end{equation}
483     where
484    
485     \begin{equation*}
486     T\text{ is absolute temperature}
487 adcroft 1.4 \end{equation*}
488 cnh 1.1 \begin{equation*}
489     p\text{ is the pressure}
490 adcroft 1.4 \end{equation*}
491 cnh 1.1 \begin{eqnarray*}
492     &&z\text{ is the height of the pressure surface} \\
493     &&g\text{ is the acceleration due to gravity}
494     \end{eqnarray*}
495    
496     In the above the ideal gas law, $p=\rho RT$, has been expressed in terms of
497     the Exner function $\Pi (p)$ given by (see Appendix Atmosphere)
498     \begin{equation}
499     \Pi (p)=c_{p}(\frac{p}{p_{c}})^{\kappa } \label{eq:exner}
500 adcroft 1.4 \end{equation}
501 cnh 1.1 where $p_{c}$ is a reference pressure and $\kappa =R/c_{p}$ with $R$ the gas
502     constant and $c_{p}$ the specific heat of air at constant pressure.
503    
504     At the top of the atmosphere (which is `fixed' in our $r$ coordinate):
505    
506     \begin{equation*}
507 cnh 1.2 R_{fixed}=p_{top}=0
508 cnh 1.1 \end{equation*}
509     In a resting atmosphere the elevation of the mountains at the bottom is
510     given by
511     \begin{equation*}
512 cnh 1.2 R_{moving}=R_{o}(x,y)=p_{o}(x,y)
513 cnh 1.1 \end{equation*}
514     i.e. the (hydrostatic) pressure at the top of the mountains in a resting
515     atmosphere.
516    
517     The boundary conditions at top and bottom are given by:
518    
519     \begin{eqnarray}
520     &&\omega =0~\text{at }r=R_{fixed} \text{ (top of the atmosphere)}
521     \label{eq:fixed-bc-atmos} \\
522     \omega &=&\frac{Dp_{s}}{Dt}\text{; at }r=R_{moving}\text{ (bottom of the
523     atmosphere)} \label{eq:moving-bc-atmos}
524     \end{eqnarray}
525    
526     Then the (hydrostatic form of) eq(\ref{eq:continuous}) yields a consistent
527     set of atmospheric equations which, for convenience, are written out in $p$
528     coordinates in Appendix Atmosphere - see eqs(\ref{eq:atmos-prime}).
529    
530     \subsection{Ocean}
531    
532     In the ocean we interpret:
533     \begin{eqnarray}
534     r &=&z\text{ is the height} \label{eq:ocean-z} \\
535     \dot{r} &=&\frac{Dz}{Dt}=w\text{ is the vertical velocity}
536     \label{eq:ocean-w} \\
537     \phi &=&\frac{p}{\rho _{c}}\text{ is the pressure} \label{eq:ocean-p} \\
538     b(\theta ,S,r) &=&\frac{g}{\rho _{c}}\left( \rho (\theta ,S,r)-\rho
539     _{c}\right) \text{ is the buoyancy} \label{eq:ocean-b}
540     \end{eqnarray}
541     where $\rho _{c}$ is a fixed reference density of water and $g$ is the
542     acceleration due to gravity.\noindent
543    
544     In the above
545    
546     At the bottom of the ocean: $R_{fixed}(x,y)=-H(x,y)$.
547    
548     The surface of the ocean is given by: $R_{moving}=\eta $
549    
550 adcroft 1.4 The position of the resting free surface of the ocean is given by $
551 cnh 1.1 R_{o}=Z_{o}=0$.
552    
553     Boundary conditions are:
554    
555     \begin{eqnarray}
556     w &=&0~\text{at }r=R_{fixed}\text{ (ocean bottom)} \label{eq:fixed-bc-ocean}
557     \\
558 adcroft 1.4 w &=&\frac{D\eta }{Dt}\text{ at }r=R_{moving}=\eta \text{ (ocean surface)
559 cnh 1.1 \label{eq:moving-bc-ocean}}
560     \end{eqnarray}
561     where $\eta $ is the elevation of the free surface.
562    
563     Then eq(\ref{eq:continuous}) yields a consistent set of oceanic equations
564     which, for convenience, are written out in $z$ coordinates in Appendix Ocean
565     - see eqs(\ref{eq:ocean-mom}) to (\ref{eq:ocean-salt}).
566    
567     \subsection{Hydrostatic, Quasi-hydrostatic, Quasi-nonhydrostatic and
568     Non-hydrostatic forms}
569    
570     Let us separate $\phi $ in to surface, hydrostatic and non-hydrostatic terms:
571    
572     \begin{equation}
573     \phi (x,y,r)=\phi _{s}(x,y)+\phi _{hyd}(x,y,r)+\phi _{nh}(x,y,r)
574     \label{eq:phi-split}
575 adcroft 1.4 \end{equation}
576 cnh 1.1 and write eq(\ref{incompressible}a,b) in the form:
577    
578     \begin{equation}
579     \frac{\partial \vec{\mathbf{v}_{h}}}{\partial t}+\mathbf{\nabla }_{h}\phi
580     _{s}+\mathbf{\nabla }_{h}\phi _{hyd}+\epsilon _{nh}\mathbf{\nabla }_{h}\phi
581     _{nh}=\vec{\mathbf{G}}_{\vec{v}_{h}} \label{eq:mom-h}
582     \end{equation}
583    
584     \begin{equation}
585     \frac{\partial \phi _{hyd}}{\partial r}=-b \label{eq:hydrostatic}
586     \end{equation}
587    
588     \begin{equation}
589 adcroft 1.4 \epsilon _{nh}\frac{\partial \dot{r}}{\partial t}+\frac{\partial \phi _{nh}}{
590 cnh 1.1 \partial r}=G_{\dot{r}} \label{eq:mom-w}
591     \end{equation}
592     Here $\epsilon _{nh}$ is a non-hydrostatic parameter.
593    
594 adcroft 1.4 The $\left( \vec{\mathbf{G}}_{\vec{v}},G_{\dot{r}}\right) $ in eq(\ref
595 cnh 1.1 {eq:mom-h}) and (\ref{eq:mom-w}) represent advective, metric and Coriolis
596 adcroft 1.4 terms in the momentum equations. In spherical coordinates they take the form
597     \footnote{
598 cnh 1.1 In the hydrostatic primitive equations (\textbf{HPE}) all underlined terms
599 adcroft 1.4 in (\ref{eq:gu-speherical}), (\ref{eq:gv-spherical}) and (\ref
600 cnh 1.1 {eq:gw-spherical}) are omitted; the singly-underlined terms are included in
601 adcroft 1.4 the quasi-hydrostatic model (\textbf{QH}). The fully non-hydrostatic model (
602 cnh 1.1 \textbf{NH}) includes all terms.} - see Marshall et al 1997a for a full
603     discussion:
604    
605     \begin{equation}
606     \left.
607     \begin{tabular}{l}
608     $G_{u}=-\vec{\mathbf{v}}.\nabla u$ \\
609 cnh 1.6 $-\left\{ \underline{\frac{u\dot{r}}{{r}}}-\frac{uv\tan \varphi}{{r}}\right\} $
610 cnh 1.1 \\
611 cnh 1.6 $-\left\{ -2\Omega v\sin \varphi+\underline{2\Omega \dot{r}\cos \varphi}\right\} $
612 cnh 1.1 \\
613 adcroft 1.4 $+\mathcal{F}_{u}$
614     \end{tabular}
615 cnh 1.1 \ \right\} \left\{
616     \begin{tabular}{l}
617     \textit{advection} \\
618     \textit{metric} \\
619     \textit{Coriolis} \\
620 adcroft 1.4 \textit{\ Forcing/Dissipation}
621     \end{tabular}
622 cnh 1.2 \ \right. \qquad \label{eq:gu-speherical}
623 cnh 1.1 \end{equation}
624    
625     \begin{equation}
626     \left.
627     \begin{tabular}{l}
628     $G_{v}=-\vec{\mathbf{v}}.\nabla v$ \\
629 cnh 1.6 $-\left\{ \underline{\frac{v\dot{r}}{{r}}}-\frac{u^{2}\tan \varphi}{{r}}\right\}
630 cnh 1.1 $ \\
631 cnh 1.6 $-\left\{ -2\Omega u\sin \varphi \right\} $ \\
632 adcroft 1.4 $+\mathcal{F}_{v}$
633     \end{tabular}
634 cnh 1.1 \ \right\} \left\{
635     \begin{tabular}{l}
636     \textit{advection} \\
637     \textit{metric} \\
638     \textit{Coriolis} \\
639 adcroft 1.4 \textit{\ Forcing/Dissipation}
640     \end{tabular}
641 cnh 1.2 \ \right. \qquad \label{eq:gv-spherical}
642 adcroft 1.4 \end{equation}
643 cnh 1.2 \qquad \qquad \qquad \qquad \qquad
644 cnh 1.1
645     \begin{equation}
646     \left.
647     \begin{tabular}{l}
648     $G_{\dot{r}}=-\underline{\underline{\vec{\mathbf{v}}.\nabla \dot{r}}}$ \\
649     $+\left\{ \underline{\frac{u^{_{^{2}}}+v^{2}}{{r}}}\right\} $ \\
650 cnh 1.6 ${+}\underline{{2\Omega u\cos \varphi}}$ \\
651 adcroft 1.4 $\underline{\underline{\mathcal{F}_{\dot{r}}}}$
652     \end{tabular}
653 cnh 1.1 \ \right\} \left\{
654     \begin{tabular}{l}
655     \textit{advection} \\
656     \textit{metric} \\
657     \textit{Coriolis} \\
658 adcroft 1.4 \textit{\ Forcing/Dissipation}
659     \end{tabular}
660 cnh 1.2 \ \right. \label{eq:gw-spherical}
661 adcroft 1.4 \end{equation}
662 cnh 1.2 \qquad \qquad \qquad \qquad \qquad
663 cnh 1.1
664 cnh 1.6 In the above `${r}$' is the distance from the center of the earth and `$\varphi$
665 cnh 1.1 ' is latitude.
666    
667     Grad and div operators in spherical coordinates are defined in appendix
668 adcroft 1.4 OPERATORS.
669 cnh 1.1 \marginpar{
670     Fig.6 Spherical polar coordinate system.}
671    
672 cnh 1.3 %%CNHbegin
673     \input{part1/sphere_coord_figure.tex}
674     %%CNHend
675    
676 cnh 1.1 \subsubsection{Shallow atmosphere approximation}
677    
678     Most models are based on the `hydrostatic primitive equations' (HPE's) in
679     which the vertical momentum equation is reduced to a statement of
680     hydrostatic balance and the `traditional approximation' is made in which the
681     Coriolis force is treated approximately and the shallow atmosphere
682     approximation is made.\ The MITgcm need not make the `traditional
683     approximation'. To be able to support consistent non-hydrostatic forms the
684 adcroft 1.4 shallow atmosphere approximation can be relaxed - when dividing through by $
685 cnh 1.2 r $ in, for example, (\ref{eq:gu-speherical}), we do not replace $r$ by $a$,
686 cnh 1.1 the radius of the earth.
687    
688     \subsubsection{Hydrostatic and quasi-hydrostatic forms}
689 cnh 1.7 \label{sec:hydrostatic_and_quasi-hydrostatic_forms}
690 cnh 1.1
691     These are discussed at length in Marshall et al (1997a).
692    
693     In the `hydrostatic primitive equations' (\textbf{HPE)} all the underlined
694     terms in Eqs. (\ref{eq:gu-speherical} $\rightarrow $\ \ref{eq:gw-spherical})
695     are neglected and `${r}$' is replaced by `$a$', the mean radius of the
696     earth. Once the pressure is found at one level - e.g. by inverting a 2-d
697     Elliptic equation for $\phi _{s}$ at $r=R_{moving}$ - the pressure can be
698 adcroft 1.4 computed at all other levels by integration of the hydrostatic relation, eq(
699 cnh 1.1 \ref{eq:hydrostatic}).
700    
701     In the `quasi-hydrostatic' equations (\textbf{QH)} strict balance between
702     gravity and vertical pressure gradients is not imposed. The $2\Omega u\cos
703 cnh 1.6 \varphi $ Coriolis term are not neglected and are balanced by a non-hydrostatic
704 adcroft 1.4 contribution to the pressure field: only the terms underlined twice in Eqs. (
705 cnh 1.1 \ref{eq:gu-speherical}$\rightarrow $\ \ref{eq:gw-spherical}) are set to zero
706     and, simultaneously, the shallow atmosphere approximation is relaxed. In
707     \textbf{QH}\ \textit{all} the metric terms are retained and the full
708     variation of the radial position of a particle monitored. The \textbf{QH}\
709     vertical momentum equation (\ref{eq:mom-w}) becomes:
710    
711     \begin{equation*}
712 cnh 1.6 \frac{\partial \phi _{nh}}{\partial r}=2\Omega u\cos \varphi
713 cnh 1.1 \end{equation*}
714     making a small correction to the hydrostatic pressure.
715    
716     \textbf{QH} has good energetic credentials - they are the same as for
717     \textbf{HPE}. Importantly, however, it has the same angular momentum
718     principle as the full non-hydrostatic model (\textbf{NH)} - see Marshall
719     et.al., 1997a. As in \textbf{HPE }only a 2-d elliptic problem need be solved.
720    
721     \subsubsection{Non-hydrostatic and quasi-nonhydrostatic forms}
722    
723     The MIT model presently supports a full non-hydrostatic ocean isomorph, but
724     only a quasi-non-hydrostatic atmospheric isomorph.
725    
726     \paragraph{Non-hydrostatic Ocean}
727    
728 adcroft 1.4 In the non-hydrostatic ocean model all terms in equations Eqs.(\ref
729 cnh 1.1 {eq:gu-speherical} $\rightarrow $\ \ref{eq:gw-spherical}) are retained. A
730     three dimensional elliptic equation must be solved subject to Neumann
731     boundary conditions (see below). It is important to note that use of the
732     full \textbf{NH} does not admit any new `fast' waves in to the system - the
733     incompressible condition eq(\ref{eq:continuous})c has already filtered out
734     acoustic modes. It does, however, ensure that the gravity waves are treated
735     accurately with an exact dispersion relation. The \textbf{NH} set has a
736     complete angular momentum principle and consistent energetics - see White
737     and Bromley, 1995; Marshall et.al.\ 1997a.
738    
739     \paragraph{Quasi-nonhydrostatic Atmosphere}
740    
741 adcroft 1.4 In the non-hydrostatic version of our atmospheric model we approximate $\dot{
742 cnh 1.1 r}$ in the vertical momentum eqs(\ref{eq:mom-w}) and (\ref{eq:gv-spherical})
743     (but only here) by:
744    
745     \begin{equation}
746     \dot{r}=\frac{Dp}{Dt}=\frac{1}{g}\frac{D\phi }{Dt} \label{eq:quasi-nh-w}
747 adcroft 1.4 \end{equation}
748 cnh 1.1 where $p_{hy}$ is the hydrostatic pressure.
749    
750     \subsubsection{Summary of equation sets supported by model}
751    
752     \paragraph{Atmosphere}
753    
754     Hydrostatic, and quasi-hydrostatic and quasi non-hydrostatic forms of the
755     compressible non-Boussinesq equations in $p-$coordinates are supported.
756    
757     \subparagraph{Hydrostatic and quasi-hydrostatic}
758    
759     The hydrostatic set is written out in $p-$coordinates in appendix Atmosphere
760     - see eq(\ref{eq:atmos-prime}).
761    
762     \subparagraph{Quasi-nonhydrostatic}
763    
764     A quasi-nonhydrostatic form is also supported.
765    
766     \paragraph{Ocean}
767    
768     \subparagraph{Hydrostatic and quasi-hydrostatic}
769    
770     Hydrostatic, and quasi-hydrostatic forms of the incompressible Boussinesq
771     equations in $z-$coordinates are supported.
772    
773     \subparagraph{Non-hydrostatic}
774    
775 adcroft 1.4 Non-hydrostatic forms of the incompressible Boussinesq equations in $z-$
776     coordinates are supported - see eqs(\ref{eq:ocean-mom}) to (\ref
777 cnh 1.1 {eq:ocean-salt}).
778    
779     \subsection{Solution strategy}
780    
781 adcroft 1.4 The method of solution employed in the \textbf{HPE}, \textbf{QH} and \textbf{
782     NH} models is summarized in Fig.7.
783 cnh 1.1 \marginpar{
784     Fig.7 Solution strategy} Under all dynamics, a 2-d elliptic equation is
785     first solved to find the surface pressure and the hydrostatic pressure at
786     any level computed from the weight of fluid above. Under \textbf{HPE} and
787     \textbf{QH} dynamics, the horizontal momentum equations are then stepped
788     forward and $\dot{r}$ found from continuity. Under \textbf{NH} dynamics a
789     3-d elliptic equation must be solved for the non-hydrostatic pressure before
790     stepping forward the horizontal momentum equations; $\dot{r}$ is found by
791     stepping forward the vertical momentum equation.
792    
793 cnh 1.3 %%CNHbegin
794     \input{part1/solution_strategy_figure.tex}
795     %%CNHend
796    
797 cnh 1.1 There is no penalty in implementing \textbf{QH} over \textbf{HPE} except, of
798 cnh 1.6 course, some complication that goes with the inclusion of $\cos \varphi \ $
799 cnh 1.1 Coriolis terms and the relaxation of the shallow atmosphere approximation.
800     But this leads to negligible increase in computation. In \textbf{NH}, in
801     contrast, one additional elliptic equation - a three-dimensional one - must
802     be inverted for $p_{nh}$. However the `overhead' of the \textbf{NH} model is
803     essentially negligible in the hydrostatic limit (see detailed discussion in
804     Marshall et al, 1997) resulting in a non-hydrostatic algorithm that, in the
805     hydrostatic limit, is as computationally economic as the \textbf{HPEs}.
806    
807     \subsection{Finding the pressure field}
808 cnh 1.7 \label{sec:finding_the_pressure_field}
809 cnh 1.1
810     Unlike the prognostic variables $u$, $v$, $w$, $\theta $ and $S$, the
811     pressure field must be obtained diagnostically. We proceed, as before, by
812     dividing the total (pressure/geo) potential in to three parts, a surface
813     part, $\phi _{s}(x,y)$, a hydrostatic part $\phi _{hyd}(x,y,r)$ and a
814     non-hydrostatic part $\phi _{nh}(x,y,r)$, as in (\ref{eq:phi-split}), and
815     writing the momentum equation as in (\ref{eq:mom-h}).
816    
817     \subsubsection{Hydrostatic pressure}
818    
819     Hydrostatic pressure is obtained by integrating (\ref{eq:hydrostatic})
820     vertically from $r=R_{o}$ where $\phi _{hyd}(r=R_{o})=0$, to yield:
821    
822     \begin{equation*}
823 adcroft 1.4 \int_{r}^{R_{o}}\frac{\partial \phi _{hyd}}{\partial r}dr=\left[ \phi _{hyd}
824 cnh 1.2 \right] _{r}^{R_{o}}=\int_{r}^{R_{o}}-bdr
825 cnh 1.1 \end{equation*}
826     and so
827    
828     \begin{equation}
829     \phi _{hyd}(x,y,r)=\int_{r}^{R_{o}}bdr \label{eq:hydro-phi}
830     \end{equation}
831    
832     The model can be easily modified to accommodate a loading term (e.g
833     atmospheric pressure pushing down on the ocean's surface) by setting:
834    
835     \begin{equation}
836     \phi _{hyd}(r=R_{o})=loading \label{eq:loading}
837     \end{equation}
838    
839     \subsubsection{Surface pressure}
840    
841 adcroft 1.4 The surface pressure equation can be obtained by integrating continuity, (
842 cnh 1.1 \ref{eq:continuous})c, vertically from $r=R_{fixed}$ to $r=R_{moving}$
843    
844     \begin{equation*}
845 adcroft 1.4 \int_{R_{fixed}}^{R_{moving}}\left( \mathbf{\nabla }_{h}\cdot \vec{\mathbf{v}
846 cnh 1.2 }_{h}+\partial _{r}\dot{r}\right) dr=0
847 cnh 1.1 \end{equation*}
848    
849     Thus:
850    
851     \begin{equation*}
852     \frac{\partial \eta }{\partial t}+\vec{\mathbf{v}}.\nabla \eta
853 adcroft 1.4 +\int_{R_{fixed}}^{R_{moving}}\mathbf{\nabla }_{h}\cdot \vec{\mathbf{v}}
854 cnh 1.2 _{h}dr=0
855 cnh 1.1 \end{equation*}
856 adcroft 1.4 where $\eta =R_{moving}-R_{o}$ is the free-surface $r$-anomaly in units of $
857 cnh 1.1 r $. The above can be rearranged to yield, using Leibnitz's theorem:
858    
859     \begin{equation}
860     \frac{\partial \eta }{\partial t}+\mathbf{\nabla }_{h}\cdot
861     \int_{R_{fixed}}^{R_{moving}}\vec{\mathbf{v}}_{h}dr=\text{source}
862     \label{eq:free-surface}
863 adcroft 1.4 \end{equation}
864 cnh 1.1 where we have incorporated a source term.
865    
866     Whether $\phi $ is pressure (ocean model, $p/\rho _{c}$) or geopotential
867     (atmospheric model), in (\ref{mtm-split}), the horizontal gradient term can
868     be written
869     \begin{equation}
870 cnh 1.2 \mathbf{\nabla }_{h}\phi _{s}=\mathbf{\nabla }_{h}\left( b_{s}\eta \right)
871 cnh 1.1 \label{eq:phi-surf}
872 adcroft 1.4 \end{equation}
873 cnh 1.1 where $b_{s}$ is the buoyancy at the surface.
874    
875 adcroft 1.4 In the hydrostatic limit ($\epsilon _{nh}=0$), Eqs(\ref{eq:mom-h}), (\ref
876 cnh 1.1 {eq:free-surface}) and (\ref{eq:phi-surf}) can be solved by inverting a 2-d
877     elliptic equation for $\phi _{s}$ as described in Chapter 2. Both `free
878     surface' and `rigid lid' approaches are available.
879    
880     \subsubsection{Non-hydrostatic pressure}
881    
882 adcroft 1.4 Taking the horizontal divergence of (\ref{hor-mtm}) and adding $\frac{
883 cnh 1.1 \partial }{\partial r}$ of (\ref{vertmtm}), invoking the continuity equation
884     (\ref{incompressible}), we deduce that:
885    
886     \begin{equation}
887 adcroft 1.4 \nabla _{3}^{2}\phi _{nh}=\nabla .\vec{\mathbf{G}}_{\vec{v}}-\left( \mathbf{
888     \nabla }_{h}^{2}\phi _{s}+\mathbf{\nabla }^{2}\phi _{hyd}\right) =\nabla .
889 cnh 1.1 \vec{\mathbf{F}} \label{eq:3d-invert}
890     \end{equation}
891    
892     For a given rhs this 3-d elliptic equation must be inverted for $\phi _{nh}$
893     subject to appropriate choice of boundary conditions. This method is usually
894     called \textit{The Pressure Method} [Harlow and Welch, 1965; Williams, 1969;
895     Potter, 1976]. In the hydrostatic primitive equations case (\textbf{HPE}),
896     the 3-d problem does not need to be solved.
897    
898     \paragraph{Boundary Conditions}
899    
900     We apply the condition of no normal flow through all solid boundaries - the
901     coasts (in the ocean) and the bottom:
902    
903     \begin{equation}
904     \vec{\mathbf{v}}.\widehat{n}=0 \label{nonormalflow}
905     \end{equation}
906     where $\widehat{n}$ is a vector of unit length normal to the boundary. The
907     kinematic condition (\ref{nonormalflow}) is also applied to the vertical
908 adcroft 1.4 velocity at $r=R_{moving}$. No-slip $\left( v_{T}=0\right) \ $or slip $
909 cnh 1.1 \left( \partial v_{T}/\partial n=0\right) \ $conditions are employed on the
910     tangential component of velocity, $v_{T}$, at all solid boundaries,
911     depending on the form chosen for the dissipative terms in the momentum
912     equations - see below.
913    
914     Eq.(\ref{nonormalflow}) implies, making use of (\ref{mtm-split}), that:
915    
916     \begin{equation}
917     \widehat{n}.\nabla \phi _{nh}=\widehat{n}.\vec{\mathbf{F}}
918     \label{eq:inhom-neumann-nh}
919     \end{equation}
920     where
921    
922     \begin{equation*}
923     \vec{\mathbf{F}}=\vec{\mathbf{G}}_{\vec{v}}-\left( \mathbf{\nabla }_{h}\phi
924     _{s}+\mathbf{\nabla }\phi _{hyd}\right)
925 adcroft 1.4 \end{equation*}
926 cnh 1.1 presenting inhomogeneous Neumann boundary conditions to the Elliptic problem
927     (\ref{eq:3d-invert}). As shown, for example, by Williams (1969), one can
928     exploit classical 3D potential theory and, by introducing an appropriately
929 cnh 1.2 chosen $\delta $-function sheet of `source-charge', replace the
930     inhomogeneous boundary condition on pressure by a homogeneous one. The
931 adcroft 1.4 source term $rhs$ in (\ref{eq:3d-invert}) is the divergence of the vector $
932     \vec{\mathbf{F}}.$ By simultaneously setting $
933 cnh 1.1 \begin{array}{l}
934 adcroft 1.4 \widehat{n}.\vec{\mathbf{F}}
935     \end{array}
936 cnh 1.1 =0$\ and $\widehat{n}.\nabla \phi _{nh}=0\ $on the boundary the following
937 cnh 1.2 self-consistent but simpler homogenized Elliptic problem is obtained:
938 cnh 1.1
939     \begin{equation*}
940 cnh 1.2 \nabla ^{2}\phi _{nh}=\nabla .\widetilde{\vec{\mathbf{F}}}\qquad
941 adcroft 1.4 \end{equation*}
942 cnh 1.1 where $\widetilde{\vec{\mathbf{F}}}$ is a modified $\vec{\mathbf{F}}$ such
943 adcroft 1.4 that $\widetilde{\vec{\mathbf{F}}}.\widehat{n}=0$. As is implied by (\ref
944 cnh 1.1 {eq:inhom-neumann-nh}) the modified boundary condition becomes:
945    
946     \begin{equation}
947     \widehat{n}.\nabla \phi _{nh}=0 \label{eq:hom-neumann-nh}
948     \end{equation}
949    
950     If the flow is `close' to hydrostatic balance then the 3-d inversion
951     converges rapidly because $\phi _{nh}\ $is then only a small correction to
952     the hydrostatic pressure field (see the discussion in Marshall et al, a,b).
953    
954     The solution $\phi _{nh}\ $to (\ref{eq:3d-invert}) and (\ref{homneuman})
955     does not vanish at $r=R_{moving}$, and so refines the pressure there.
956    
957     \subsection{Forcing/dissipation}
958    
959     \subsubsection{Forcing}
960    
961     The forcing terms $\mathcal{F}$ on the rhs of the equations are provided by
962     `physics packages' described in detail in chapter ??.
963    
964     \subsubsection{Dissipation}
965    
966     \paragraph{Momentum}
967    
968     Many forms of momentum dissipation are available in the model. Laplacian and
969     biharmonic frictions are commonly used:
970    
971     \begin{equation}
972 adcroft 1.4 D_{V}=A_{h}\nabla _{h}^{2}v+A_{v}\frac{\partial ^{2}v}{\partial z^{2}}
973 cnh 1.1 +A_{4}\nabla _{h}^{4}v \label{eq:dissipation}
974     \end{equation}
975     where $A_{h}$ and $A_{v}\ $are (constant) horizontal and vertical viscosity
976     coefficients and $A_{4}\ $is the horizontal coefficient for biharmonic
977     friction. These coefficients are the same for all velocity components.
978    
979     \paragraph{Tracers}
980    
981     The mixing terms for the temperature and salinity equations have a similar
982     form to that of momentum except that the diffusion tensor can be
983 adcroft 1.4 non-diagonal and have varying coefficients. $\qquad $
984 cnh 1.1 \begin{equation}
985     D_{T,S}=\nabla .[\underline{\underline{K}}\nabla (T,S)]+K_{4}\nabla
986     _{h}^{4}(T,S) \label{eq:diffusion}
987     \end{equation}
988 adcroft 1.4 where $\underline{\underline{K}}\ $is the diffusion tensor and the $K_{4}\ $
989 cnh 1.1 horizontal coefficient for biharmonic diffusion. In the simplest case where
990     the subgrid-scale fluxes of heat and salt are parameterized with constant
991     horizontal and vertical diffusion coefficients, $\underline{\underline{K}}$,
992     reduces to a diagonal matrix with constant coefficients:
993    
994     \begin{equation}
995     \qquad \qquad \qquad \qquad K=\left(
996     \begin{array}{ccc}
997     K_{h} & 0 & 0 \\
998     0 & K_{h} & 0 \\
999 adcroft 1.4 0 & 0 & K_{v}
1000 cnh 1.1 \end{array}
1001     \right) \qquad \qquad \qquad \label{eq:diagonal-diffusion-tensor}
1002     \end{equation}
1003     where $K_{h}\ $and $K_{v}\ $are the horizontal and vertical diffusion
1004     coefficients. These coefficients are the same for all tracers (temperature,
1005     salinity ... ).
1006    
1007     \subsection{Vector invariant form}
1008    
1009 adcroft 1.4 For some purposes it is advantageous to write momentum advection in eq(\ref
1010 cnh 1.1 {hor-mtm}) and (\ref{vertmtm}) in the (so-called) `vector invariant' form:
1011    
1012     \begin{equation}
1013 adcroft 1.4 \frac{D\vec{\mathbf{v}}}{Dt}=\frac{\partial \vec{\mathbf{v}}}{\partial t}
1014     +\left( \nabla \times \vec{\mathbf{v}}\right) \times \vec{\mathbf{v}}+\nabla
1015 cnh 1.2 \left[ \frac{1}{2}(\vec{\mathbf{v}}\cdot \vec{\mathbf{v}})\right]
1016 cnh 1.1 \label{eq:vi-identity}
1017 adcroft 1.4 \end{equation}
1018 cnh 1.1 This permits alternative numerical treatments of the non-linear terms based
1019     on their representation as a vorticity flux. Because gradients of coordinate
1020     vectors no longer appear on the rhs of (\ref{eq:vi-identity}), explicit
1021 adcroft 1.4 representation of the metric terms in (\ref{eq:gu-speherical}), (\ref
1022 cnh 1.1 {eq:gv-spherical}) and (\ref{eq:gw-spherical}), can be avoided: information
1023     about the geometry is contained in the areas and lengths of the volumes used
1024     to discretize the model.
1025    
1026     \subsection{Adjoint}
1027    
1028 cnh 1.2 Tangent linear and adjoint counterparts of the forward model and described
1029     in Chapter 5.
1030 cnh 1.1
1031 cnh 1.7 % $Header: /u/u0/gcmpack/mitgcmdoc/part1/manual.tex,v 1.6 2001/10/24 15:21:27 cnh Exp $
1032 cnh 1.1 % $Name: $
1033    
1034     \section{Appendix ATMOSPHERE}
1035    
1036     \subsection{Hydrostatic Primitive Equations for the Atmosphere in pressure
1037     coordinates}
1038    
1039     \label{sect-hpe-p}
1040    
1041     The hydrostatic primitive equations (HPEs) in p-coordinates are:
1042     \begin{eqnarray}
1043 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1044 cnh 1.2 _{h}+\mathbf{\nabla }_{p}\phi &=&\vec{\mathbf{\mathcal{F}}}
1045 cnh 1.1 \label{eq:atmos-mom} \\
1046 cnh 1.2 \frac{\partial \phi }{\partial p}+\alpha &=&0 \label{eq-p-hydro-start} \\
1047 adcroft 1.4 \mathbf{\nabla }_{p}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \omega }{
1048 cnh 1.1 \partial p} &=&0 \label{eq:atmos-cont} \\
1049 cnh 1.2 p\alpha &=&RT \label{eq:atmos-eos} \\
1050 cnh 1.1 c_{v}\frac{DT}{Dt}+p\frac{D\alpha }{Dt} &=&\mathcal{Q} \label{eq:atmos-heat}
1051 adcroft 1.4 \end{eqnarray}
1052 cnh 1.1 where $\vec{\mathbf{v}}_{h}=(u,v,0)$ is the `horizontal' (on pressure
1053     surfaces) component of velocity,$\frac{D}{Dt}=\vec{\mathbf{v}}_{h}\cdot
1054     \mathbf{\nabla }_{p}+\omega \frac{\partial }{\partial p}$ is the total
1055 cnh 1.6 derivative, $f=2\Omega \sin \varphi$ is the Coriolis parameter, $\phi =gz$ is
1056 adcroft 1.4 the geopotential, $\alpha =1/\rho $ is the specific volume, $\omega =\frac{Dp
1057     }{Dt}$ is the vertical velocity in the $p-$coordinate. Equation(\ref
1058     {eq:atmos-heat}) is the first law of thermodynamics where internal energy $
1059     e=c_{v}T$, $T$ is temperature, $Q$ is the rate of heating per unit mass and $
1060 cnh 1.1 p\frac{D\alpha }{Dt}$ is the work done by the fluid in compressing.
1061    
1062     It is convenient to cast the heat equation in terms of potential temperature
1063     $\theta $ so that it looks more like a generic conservation law.
1064     Differentiating (\ref{eq:atmos-eos}) we get:
1065     \begin{equation*}
1066     p\frac{D\alpha }{Dt}+\alpha \frac{Dp}{Dt}=R\frac{DT}{Dt}
1067 adcroft 1.4 \end{equation*}
1068     which, when added to the heat equation (\ref{eq:atmos-heat}) and using $
1069 cnh 1.1 c_{p}=c_{v}+R$, gives:
1070     \begin{equation}
1071     c_{p}\frac{DT}{Dt}-\alpha \frac{Dp}{Dt}=\mathcal{Q}
1072     \label{eq-p-heat-interim}
1073 adcroft 1.4 \end{equation}
1074 cnh 1.1 Potential temperature is defined:
1075     \begin{equation}
1076     \theta =T(\frac{p_{c}}{p})^{\kappa } \label{eq:potential-temp}
1077 adcroft 1.4 \end{equation}
1078 cnh 1.1 where $p_{c}$ is a reference pressure and $\kappa =R/c_{p}$. For convenience
1079     we will make use of the Exner function $\Pi (p)$ which defined by:
1080     \begin{equation}
1081     \Pi (p)=c_{p}(\frac{p}{p_{c}})^{\kappa } \label{Exner}
1082 adcroft 1.4 \end{equation}
1083 cnh 1.1 The following relations will be useful and are easily expressed in terms of
1084     the Exner function:
1085     \begin{equation*}
1086     c_{p}T=\Pi \theta \;\;;\;\;\frac{\partial \Pi }{\partial p}=\frac{\kappa \Pi
1087 adcroft 1.4 }{p}\;\;;\;\;\alpha =\frac{\kappa \Pi \theta }{p}=\frac{\partial \ \Pi }{
1088     \partial p}\theta \;\;;\;\;\frac{D\Pi }{Dt}=\frac{\partial \Pi }{\partial p}
1089 cnh 1.1 \frac{Dp}{Dt}
1090 adcroft 1.4 \end{equation*}
1091 cnh 1.1 where $b=\frac{\partial \ \Pi }{\partial p}\theta $ is the buoyancy.
1092    
1093     The heat equation is obtained by noting that
1094     \begin{equation*}
1095     c_{p}\frac{DT}{Dt}=\frac{D(\Pi \theta )}{Dt}=\Pi \frac{D\theta }{Dt}+\theta
1096 cnh 1.2 \frac{D\Pi }{Dt}=\Pi \frac{D\theta }{Dt}+\alpha \frac{Dp}{Dt}
1097 cnh 1.1 \end{equation*}
1098     and on substituting into (\ref{eq-p-heat-interim}) gives:
1099     \begin{equation}
1100     \Pi \frac{D\theta }{Dt}=\mathcal{Q}
1101     \label{eq:potential-temperature-equation}
1102     \end{equation}
1103     which is in conservative form.
1104    
1105 adcroft 1.4 For convenience in the model we prefer to step forward (\ref
1106 cnh 1.1 {eq:potential-temperature-equation}) rather than (\ref{eq:atmos-heat}).
1107    
1108     \subsubsection{Boundary conditions}
1109    
1110     The upper and lower boundary conditions are :
1111     \begin{eqnarray}
1112     \mbox{at the top:}\;\;p=0 &&\text{, }\omega =\frac{Dp}{Dt}=0 \\
1113     \mbox{at the surface:}\;\;p=p_{s} &&\text{, }\phi =\phi _{topo}=g~Z_{topo}
1114     \label{eq:boundary-condition-atmosphere}
1115     \end{eqnarray}
1116     In $p$-coordinates, the upper boundary acts like a solid boundary ($\omega
1117     =0 $); in $z$-coordinates and the lower boundary is analogous to a free
1118     surface ($\phi $ is imposed and $\omega \neq 0$).
1119    
1120     \subsubsection{Splitting the geo-potential}
1121    
1122     For the purposes of initialization and reducing round-off errors, the model
1123     deals with perturbations from reference (or ``standard'') profiles. For
1124     example, the hydrostatic geopotential associated with the resting atmosphere
1125     is not dynamically relevant and can therefore be subtracted from the
1126     equations. The equations written in terms of perturbations are obtained by
1127     substituting the following definitions into the previous model equations:
1128     \begin{eqnarray}
1129     \theta &=&\theta _{o}+\theta ^{\prime } \label{eq:atmos-ref-prof-theta} \\
1130     \alpha &=&\alpha _{o}+\alpha ^{\prime } \label{eq:atmos-ref-prof-alpha} \\
1131     \phi &=&\phi _{o}+\phi ^{\prime } \label{eq:atmos-ref-prof-phi}
1132     \end{eqnarray}
1133     The reference state (indicated by subscript ``0'') corresponds to
1134     horizontally homogeneous atmosphere at rest ($\theta _{o},\alpha _{o},\phi
1135     _{o}$) with surface pressure $p_{o}(x,y)$ that satisfies $\phi
1136     _{o}(p_{o})=g~Z_{topo}$, defined:
1137     \begin{eqnarray*}
1138     \theta _{o}(p) &=&f^{n}(p) \\
1139     \alpha _{o}(p) &=&\Pi _{p}\theta _{o} \\
1140     \phi _{o}(p) &=&\phi _{topo}-\int_{p_{0}}^{p}\alpha _{o}dp
1141     \end{eqnarray*}
1142     %\begin{eqnarray*}
1143     %\phi'_\alpha & = & \int^p_{p_o} (\alpha_o -\alpha) dp \\
1144     %\phi'_s(x,y,t) & = & \int_{p_o}^{p_s} \alpha dp
1145     %\end{eqnarray*}
1146    
1147     The final form of the HPE's in p coordinates is then:
1148     \begin{eqnarray}
1149 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1150 cnh 1.1 _{h}+\mathbf{\nabla }_{p}\phi ^{\prime } &=&\vec{\mathbf{\mathcal{F}}} \\
1151     \frac{\partial \phi ^{\prime }}{\partial p}+\alpha ^{\prime } &=&0 \\
1152 adcroft 1.4 \mathbf{\nabla }_{p}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \omega }{
1153 cnh 1.1 \partial p} &=&0 \\
1154     \frac{\partial \Pi }{\partial p}\theta ^{\prime } &=&\alpha ^{\prime } \\
1155     \frac{D\theta }{Dt} &=&\frac{\mathcal{Q}}{\Pi } \label{eq:atmos-prime}
1156     \end{eqnarray}
1157    
1158 cnh 1.7 % $Header: /u/u0/gcmpack/mitgcmdoc/part1/manual.tex,v 1.6 2001/10/24 15:21:27 cnh Exp $
1159 cnh 1.1 % $Name: $
1160    
1161     \section{Appendix OCEAN}
1162    
1163     \subsection{Equations of motion for the ocean}
1164    
1165     We review here the method by which the standard (Boussinesq, incompressible)
1166     HPE's for the ocean written in z-coordinates are obtained. The
1167     non-Boussinesq equations for oceanic motion are:
1168     \begin{eqnarray}
1169 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1170 cnh 1.1 _{h}+\frac{1}{\rho }\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}} \\
1171     \epsilon _{nh}\frac{Dw}{Dt}+g+\frac{1}{\rho }\frac{\partial p}{\partial z}
1172     &=&\epsilon _{nh}\mathcal{F}_{w} \\
1173 adcroft 1.4 \frac{1}{\rho }\frac{D\rho }{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}
1174 cnh 1.1 _{h}+\frac{\partial w}{\partial z} &=&0 \\
1175 cnh 1.2 \rho &=&\rho (\theta ,S,p) \\
1176 cnh 1.1 \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \\
1177     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq:non-boussinesq}
1178 adcroft 1.4 \end{eqnarray}
1179 cnh 1.1 These equations permit acoustics modes, inertia-gravity waves,
1180     non-hydrostatic motions, a geostrophic (Rossby) mode and a thermo-haline
1181     mode. As written, they cannot be integrated forward consistently - if we
1182     step $\rho $ forward in (\ref{eq-zns-cont}), the answer will not be
1183 adcroft 1.4 consistent with that obtained by stepping (\ref{eq-zns-heat}) and (\ref
1184 cnh 1.1 {eq-zns-salt}) and then using (\ref{eq-zns-eos}) to yield $\rho $. It is
1185     therefore necessary to manipulate the system as follows. Differentiating the
1186     EOS (equation of state) gives:
1187    
1188     \begin{equation}
1189     \frac{D\rho }{Dt}=\left. \frac{\partial \rho }{\partial \theta }\right|
1190     _{S,p}\frac{D\theta }{Dt}+\left. \frac{\partial \rho }{\partial S}\right|
1191     _{\theta ,p}\frac{DS}{Dt}+\left. \frac{\partial \rho }{\partial p}\right|
1192     _{\theta ,S}\frac{Dp}{Dt} \label{EOSexpansion}
1193     \end{equation}
1194    
1195     Note that $\frac{\partial \rho }{\partial p}=\frac{1}{c_{s}^{2}}$ is the
1196 adcroft 1.4 reciprocal of the sound speed ($c_{s}$) squared. Substituting into \ref
1197 cnh 1.1 {eq-zns-cont} gives:
1198     \begin{equation}
1199 adcroft 1.4 \frac{1}{\rho c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{
1200 cnh 1.1 v}}+\partial _{z}w\approx 0 \label{eq-zns-pressure}
1201     \end{equation}
1202     where we have used an approximation sign to indicate that we have assumed
1203     adiabatic motion, dropping the $\frac{D\theta }{Dt}$ and $\frac{DS}{Dt}$.
1204     Replacing \ref{eq-zns-cont} with \ref{eq-zns-pressure} yields a system that
1205     can be explicitly integrated forward:
1206     \begin{eqnarray}
1207 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1208 cnh 1.1 _{h}+\frac{1}{\rho }\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1209     \label{eq-cns-hmom} \\
1210     \epsilon _{nh}\frac{Dw}{Dt}+g+\frac{1}{\rho }\frac{\partial p}{\partial z}
1211     &=&\epsilon _{nh}\mathcal{F}_{w} \label{eq-cns-hydro} \\
1212 adcroft 1.4 \frac{1}{\rho c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{
1213 cnh 1.1 v}}_{h}+\frac{\partial w}{\partial z} &=&0 \label{eq-cns-cont} \\
1214     \rho &=&\rho (\theta ,S,p) \label{eq-cns-eos} \\
1215     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-cns-heat} \\
1216     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-cns-salt}
1217     \end{eqnarray}
1218    
1219     \subsubsection{Compressible z-coordinate equations}
1220    
1221     Here we linearize the acoustic modes by replacing $\rho $ with $\rho _{o}(z)$
1222     wherever it appears in a product (ie. non-linear term) - this is the
1223     `Boussinesq assumption'. The only term that then retains the full variation
1224     in $\rho $ is the gravitational acceleration:
1225     \begin{eqnarray}
1226 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1227 cnh 1.1 _{h}+\frac{1}{\rho _{o}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1228     \label{eq-zcb-hmom} \\
1229 adcroft 1.4 \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{o}}+\frac{1}{\rho _{o}}
1230 cnh 1.1 \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1231     \label{eq-zcb-hydro} \\
1232 adcroft 1.4 \frac{1}{\rho _{o}c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{
1233 cnh 1.1 \mathbf{v}}_{h}+\frac{\partial w}{\partial z} &=&0 \label{eq-zcb-cont} \\
1234     \rho &=&\rho (\theta ,S,p) \label{eq-zcb-eos} \\
1235     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-zcb-heat} \\
1236     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-zcb-salt}
1237     \end{eqnarray}
1238     These equations still retain acoustic modes. But, because the
1239 adcroft 1.4 ``compressible'' terms are linearized, the pressure equation \ref
1240 cnh 1.1 {eq-zcb-cont} can be integrated implicitly with ease (the time-dependent
1241     term appears as a Helmholtz term in the non-hydrostatic pressure equation).
1242     These are the \emph{truly} compressible Boussinesq equations. Note that the
1243     EOS must have the same pressure dependency as the linearized pressure term,
1244 adcroft 1.4 ie. $\left. \frac{\partial \rho }{\partial p}\right| _{\theta ,S}=\frac{1}{
1245 cnh 1.1 c_{s}^{2}}$, for consistency.
1246    
1247     \subsubsection{`Anelastic' z-coordinate equations}
1248    
1249     The anelastic approximation filters the acoustic mode by removing the
1250 adcroft 1.4 time-dependency in the continuity (now pressure-) equation (\ref{eq-zcb-cont}
1251     ). This could be done simply by noting that $\frac{Dp}{Dt}\approx -g\rho _{o}
1252 cnh 1.1 \frac{Dz}{Dt}=-g\rho _{o}w$, but this leads to an inconsistency between
1253     continuity and EOS. A better solution is to change the dependency on
1254     pressure in the EOS by splitting the pressure into a reference function of
1255     height and a perturbation:
1256     \begin{equation*}
1257 cnh 1.2 \rho =\rho (\theta ,S,p_{o}(z)+\epsilon _{s}p^{\prime })
1258 cnh 1.1 \end{equation*}
1259     Remembering that the term $\frac{Dp}{Dt}$ in continuity comes from
1260     differentiating the EOS, the continuity equation then becomes:
1261     \begin{equation*}
1262 adcroft 1.4 \frac{1}{\rho _{o}c_{s}^{2}}\left( \frac{Dp_{o}}{Dt}+\epsilon _{s}\frac{
1263     Dp^{\prime }}{Dt}\right) +\mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+
1264 cnh 1.2 \frac{\partial w}{\partial z}=0
1265 cnh 1.1 \end{equation*}
1266     If the time- and space-scales of the motions of interest are longer than
1267 adcroft 1.4 those of acoustic modes, then $\frac{Dp^{\prime }}{Dt}<<(\frac{Dp_{o}}{Dt},
1268 cnh 1.1 \mathbf{\nabla }\cdot \vec{\mathbf{v}}_{h})$ in the continuity equations and
1269 adcroft 1.4 $\left. \frac{\partial \rho }{\partial p}\right| _{\theta ,S}\frac{
1270 cnh 1.1 Dp^{\prime }}{Dt}<<\left. \frac{\partial \rho }{\partial p}\right| _{\theta
1271     ,S}\frac{Dp_{o}}{Dt}$ in the EOS (\ref{EOSexpansion}). Thus we set $\epsilon
1272     _{s}=0$, removing the dependency on $p^{\prime }$ in the continuity equation
1273     and EOS. Expanding $\frac{Dp_{o}(z)}{Dt}=-g\rho _{o}w$ then leads to the
1274     anelastic continuity equation:
1275     \begin{equation}
1276 adcroft 1.4 \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z}-
1277 cnh 1.1 \frac{g}{c_{s}^{2}}w=0 \label{eq-za-cont1}
1278     \end{equation}
1279     A slightly different route leads to the quasi-Boussinesq continuity equation
1280 adcroft 1.4 where we use the scaling $\frac{\partial \rho ^{\prime }}{\partial t}+
1281     \mathbf{\nabla }_{3}\cdot \rho ^{\prime }\vec{\mathbf{v}}<<\mathbf{\nabla }
1282 cnh 1.1 _{3}\cdot \rho _{o}\vec{\mathbf{v}}$ yielding:
1283     \begin{equation}
1284 adcroft 1.4 \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{1}{\rho _{o}}\frac{
1285 cnh 1.1 \partial \left( \rho _{o}w\right) }{\partial z}=0 \label{eq-za-cont2}
1286     \end{equation}
1287     Equations \ref{eq-za-cont1} and \ref{eq-za-cont2} are in fact the same
1288     equation if:
1289     \begin{equation}
1290     \frac{1}{\rho _{o}}\frac{\partial \rho _{o}}{\partial z}=\frac{-g}{c_{s}^{2}}
1291     \end{equation}
1292     Again, note that if $\rho _{o}$ is evaluated from prescribed $\theta _{o}$
1293 adcroft 1.4 and $S_{o}$ profiles, then the EOS dependency on $p_{o}$ and the term $\frac{
1294 cnh 1.1 g}{c_{s}^{2}}$ in continuity should be referred to those same profiles. The
1295     full set of `quasi-Boussinesq' or `anelastic' equations for the ocean are
1296     then:
1297     \begin{eqnarray}
1298 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1299 cnh 1.1 _{h}+\frac{1}{\rho _{o}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1300     \label{eq-zab-hmom} \\
1301 adcroft 1.4 \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{o}}+\frac{1}{\rho _{o}}
1302 cnh 1.1 \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1303     \label{eq-zab-hydro} \\
1304 adcroft 1.4 \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{1}{\rho _{o}}\frac{
1305 cnh 1.1 \partial \left( \rho _{o}w\right) }{\partial z} &=&0 \label{eq-zab-cont} \\
1306     \rho &=&\rho (\theta ,S,p_{o}(z)) \label{eq-zab-eos} \\
1307     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-zab-heat} \\
1308     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-zab-salt}
1309     \end{eqnarray}
1310    
1311     \subsubsection{Incompressible z-coordinate equations}
1312    
1313     Here, the objective is to drop the depth dependence of $\rho _{o}$ and so,
1314     technically, to also remove the dependence of $\rho $ on $p_{o}$. This would
1315     yield the ``truly'' incompressible Boussinesq equations:
1316     \begin{eqnarray}
1317 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1318 cnh 1.1 _{h}+\frac{1}{\rho _{c}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1319     \label{eq-ztb-hmom} \\
1320 adcroft 1.4 \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{c}}+\frac{1}{\rho _{c}}
1321 cnh 1.1 \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1322     \label{eq-ztb-hydro} \\
1323     \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z}
1324     &=&0 \label{eq-ztb-cont} \\
1325     \rho &=&\rho (\theta ,S) \label{eq-ztb-eos} \\
1326     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-ztb-heat} \\
1327     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-ztb-salt}
1328     \end{eqnarray}
1329     where $\rho _{c}$ is a constant reference density of water.
1330    
1331     \subsubsection{Compressible non-divergent equations}
1332    
1333     The above ``incompressible'' equations are incompressible in both the flow
1334     and the density. In many oceanic applications, however, it is important to
1335     retain compressibility effects in the density. To do this we must split the
1336     density thus:
1337     \begin{equation*}
1338     \rho =\rho _{o}+\rho ^{\prime }
1339 adcroft 1.4 \end{equation*}
1340 cnh 1.1 We then assert that variations with depth of $\rho _{o}$ are unimportant
1341     while the compressible effects in $\rho ^{\prime }$ are:
1342     \begin{equation*}
1343     \rho _{o}=\rho _{c}
1344 adcroft 1.4 \end{equation*}
1345 cnh 1.1 \begin{equation*}
1346     \rho ^{\prime }=\rho (\theta ,S,p_{o}(z))-\rho _{o}
1347 adcroft 1.4 \end{equation*}
1348 cnh 1.1 This then yields what we can call the semi-compressible Boussinesq
1349     equations:
1350     \begin{eqnarray}
1351 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1352     _{h}+\frac{1}{\rho _{c}}\mathbf{\nabla }_{z}p^{\prime } &=&\vec{\mathbf{
1353 cnh 1.1 \mathcal{F}}} \label{eq:ocean-mom} \\
1354     \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho ^{\prime }}{\rho _{c}}+\frac{1}{\rho
1355     _{c}}\frac{\partial p^{\prime }}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1356     \label{eq:ocean-wmom} \\
1357     \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z}
1358     &=&0 \label{eq:ocean-cont} \\
1359     \rho ^{\prime } &=&\rho (\theta ,S,p_{o}(z))-\rho _{c} \label{eq:ocean-eos}
1360     \\
1361     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq:ocean-theta} \\
1362     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq:ocean-salt}
1363 adcroft 1.4 \end{eqnarray}
1364 cnh 1.1 Note that the hydrostatic pressure of the resting fluid, including that
1365     associated with $\rho _{c}$, is subtracted out since it has no effect on the
1366     dynamics.
1367    
1368     Though necessary, the assumptions that go into these equations are messy
1369     since we essentially assume a different EOS for the reference density and
1370     the perturbation density. Nevertheless, it is the hydrostatic ($\epsilon
1371     _{nh}=0$ form of these equations that are used throughout the ocean modeling
1372     community and referred to as the primitive equations (HPE).
1373    
1374 cnh 1.7 % $Header: /u/u0/gcmpack/mitgcmdoc/part1/manual.tex,v 1.6 2001/10/24 15:21:27 cnh Exp $
1375 cnh 1.1 % $Name: $
1376    
1377     \section{Appendix:OPERATORS}
1378    
1379     \subsection{Coordinate systems}
1380    
1381     \subsubsection{Spherical coordinates}
1382    
1383     In spherical coordinates, the velocity components in the zonal, meridional
1384     and vertical direction respectively, are given by (see Fig.2) :
1385    
1386     \begin{equation*}
1387 cnh 1.6 u=r\cos \varphi \frac{D\lambda }{Dt}
1388 cnh 1.1 \end{equation*}
1389    
1390     \begin{equation*}
1391 cnh 1.6 v=r\frac{D\varphi }{Dt}\qquad
1392 cnh 1.1 \end{equation*}
1393     $\qquad \qquad \qquad \qquad $
1394    
1395     \begin{equation*}
1396 cnh 1.2 \dot{r}=\frac{Dr}{Dt}
1397 cnh 1.1 \end{equation*}
1398    
1399 cnh 1.6 Here $\varphi $ is the latitude, $\lambda $ the longitude, $r$ the radial
1400 cnh 1.1 distance of the particle from the center of the earth, $\Omega $ is the
1401     angular speed of rotation of the Earth and $D/Dt$ is the total derivative.
1402    
1403     The `grad' ($\nabla $) and `div' ($\nabla $.) operators are defined by, in
1404     spherical coordinates:
1405    
1406     \begin{equation*}
1407 cnh 1.6 \nabla \equiv \left( \frac{1}{r\cos \varphi }\frac{\partial }{\partial \lambda }
1408     ,\frac{1}{r}\frac{\partial }{\partial \varphi },\frac{\partial }{\partial r}
1409 cnh 1.2 \right)
1410 cnh 1.1 \end{equation*}
1411    
1412     \begin{equation*}
1413 cnh 1.6 \nabla .v\equiv \frac{1}{r\cos \varphi }\left\{ \frac{\partial u}{\partial
1414     \lambda }+\frac{\partial }{\partial \varphi }\left( v\cos \varphi \right) \right\}
1415 cnh 1.2 +\frac{1}{r^{2}}\frac{\partial \left( r^{2}\dot{r}\right) }{\partial r}
1416 cnh 1.1 \end{equation*}
1417    
1418 adcroft 1.4 %tci%\end{document}

  ViewVC Help
Powered by ViewVC 1.1.22