/[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 1.13 - (hide annotations) (download) (as text)
Mon Nov 19 19:58:20 2001 UTC (23 years, 7 months ago) by cnh
Branch: MAIN
Changes since 1.12: +8 -8 lines
File MIME type: application/x-tex
Fig. layouts

1 cnh 1.13 % $Header: /u/u0/gcmpack/mitgcmdoc/part1/manual.tex,v 1.12 2001/11/19 14:33:44 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 1.13 % $Header: /u/u0/gcmpack/mitgcmdoc/part1/manual.tex,v 1.12 2001/11/19 14:33:44 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 cnh 1.12 Key publications reporting on and charting the development of the model are:
87    
88     \begin{verbatim}
89    
90     Hill, C. and J. Marshall, (1995)
91     Application of a Parallel Navier-Stokes Model to Ocean Circulation in
92     Parallel Computational Fluid Dynamics
93     In Proceedings of Parallel Computational Fluid Dynamics: Implementations
94     and Results Using Parallel Computers, 545-552.
95     Elsevier Science B.V.: New York
96    
97     Marshall, J., C. Hill, L. Perelman, and A. Adcroft, (1997)
98     Hydrostatic, quasi-hydrostatic, and nonhydrostatic ocean modeling,
99     J. Geophysical Res., 102(C3), 5733-5752.
100    
101     Marshall, J., A. Adcroft, C. Hill, L. Perelman, and C. Heisey, (1997)
102     A finite-volume, incompressible Navier Stokes model for studies of the ocean
103     on parallel computers,
104     J. Geophysical Res., 102(C3), 5753-5766.
105    
106     Adcroft, A.J., Hill, C.N. and J. Marshall, (1997)
107     Representation of topography by shaved cells in a height coordinate ocean
108     model
109     Mon Wea Rev, vol 125, 2293-2315
110    
111     Marshall, J., Jones, H. and C. Hill, (1998)
112     Efficient ocean modeling using non-hydrostatic algorithms
113     Journal of Marine Systems, 18, 115-134
114    
115     Adcroft, A., Hill C. and J. Marshall: (1999)
116     A new treatment of the Coriolis terms in C-grid models at both high and low
117     resolutions,
118     Mon. Wea. Rev. Vol 127, pages 1928-1936
119    
120     Hill, C, Adcroft,A., Jamous,D., and J. Marshall, (1999)
121     A Strategy for Terascale Climate Modeling.
122     In Proceedings of the Eight ECMWF Workshop on the Use of Parallel Processors
123     in Meteorology
124    
125     Marotzke, J, Giering,R., Zhang, K.Q., Stammer,D., Hill,C., and T.Lee, (1999)
126     Construction of the adjoint MIT ocean general circulation model and
127     application to Atlantic heat transport variability
128     J. Geophysical Res., 104(C12), 29,529-29,547.
129    
130    
131     \end{verbatim}
132 cnh 1.1
133     We begin by briefly showing some of the results of the model in action to
134     give a feel for the wide range of problems that can be addressed using it.
135    
136 cnh 1.13 % $Header: /u/u0/gcmpack/mitgcmdoc/part1/manual.tex,v 1.12 2001/11/19 14:33:44 cnh Exp $
137 cnh 1.1 % $Name: $
138    
139     \section{Illustrations of the model in action}
140    
141     The MITgcm has been designed and used to model a wide range of phenomena,
142     from convection on the scale of meters in the ocean to the global pattern of
143 cnh 1.7 atmospheric winds - see figure \ref{fig:all-scales}. To give a flavor of the
144 cnh 1.1 kinds of problems the model has been used to study, we briefly describe some
145     of them here. A more detailed description of the underlying formulation,
146     numerical algorithm and implementation that lie behind these calculations is
147 cnh 1.2 given later. Indeed many of the illustrative examples shown below can be
148     easily reproduced: simply download the model (the minimum you need is a PC
149 cnh 1.10 running Linux, together with a FORTRAN\ 77 compiler) and follow the examples
150 cnh 1.2 described in detail in the documentation.
151 cnh 1.1
152     \subsection{Global atmosphere: `Held-Suarez' benchmark}
153    
154 cnh 1.7 A novel feature of MITgcm is its ability to simulate, using one basic algorithm,
155     both atmospheric and oceanographic flows at both small and large scales.
156 cnh 1.2
157 cnh 1.7 Figure \ref{fig:eddy_cs} shows an instantaneous plot of the 500$mb$
158 cnh 1.2 temperature field obtained using the atmospheric isomorph of MITgcm run at
159     2.8$^{\circ }$ resolution on the cubed sphere. We see cold air over the pole
160     (blue) and warm air along an equatorial band (red). Fully developed
161     baroclinic eddies spawned in the northern hemisphere storm track are
162     evident. There are no mountains or land-sea contrast in this calculation,
163     but you can easily put them in. The model is driven by relaxation to a
164     radiative-convective equilibrium profile, following the description set out
165     in Held and Suarez; 1994 designed to test atmospheric hydrodynamical cores -
166     there are no mountains or land-sea contrast.
167    
168 cnh 1.3 %% CNHbegin
169     \input{part1/cubic_eddies_figure}
170     %% CNHend
171    
172 cnh 1.2 As described in Adcroft (2001), a `cubed sphere' is used to discretize the
173 cnh 1.10 globe permitting a uniform griding and obviated the need to Fourier filter.
174 cnh 1.2 The `vector-invariant' form of MITgcm supports any orthogonal curvilinear
175     grid, of which the cubed sphere is just one of many choices.
176 cnh 1.1
177 cnh 1.7 Figure \ref{fig:hs_zave_u} shows the 5-year mean, zonally averaged zonal
178     wind from a 20-level configuration of
179 cnh 1.2 the model. It compares favorable with more conventional spatial
180 cnh 1.7 discretization approaches. The two plots show the field calculated using the
181     cube-sphere grid and the flow calculated using a regular, spherical polar
182     latitude-longitude grid. Both grids are supported within the model.
183 cnh 1.1
184 cnh 1.3 %% CNHbegin
185     \input{part1/hs_zave_u_figure}
186     %% CNHend
187    
188 cnh 1.2 \subsection{Ocean gyres}
189 cnh 1.1
190 cnh 1.2 Baroclinic instability is a ubiquitous process in the ocean, as well as the
191     atmosphere. Ocean eddies play an important role in modifying the
192     hydrographic structure and current systems of the oceans. Coarse resolution
193     models of the oceans cannot resolve the eddy field and yield rather broad,
194     diffusive patterns of ocean currents. But if the resolution of our models is
195     increased until the baroclinic instability process is resolved, numerical
196     solutions of a different and much more realistic kind, can be obtained.
197    
198 cnh 1.7 Figure \ref{fig:ocean-gyres} shows the surface temperature and velocity
199     field obtained from MITgcm run at $\frac{1}{6}^{\circ }$ horizontal
200     resolution on a $lat-lon$
201 cnh 1.2 grid in which the pole has been rotated by 90$^{\circ }$ on to the equator
202     (to avoid the converging of meridian in northern latitudes). 21 vertical
203     levels are used in the vertical with a `lopped cell' representation of
204     topography. The development and propagation of anomalously warm and cold
205 cnh 1.7 eddies can be clearly seen in the Gulf Stream region. The transport of
206 cnh 1.2 warm water northward by the mean flow of the Gulf Stream is also clearly
207     visible.
208 cnh 1.1
209 cnh 1.3 %% CNHbegin
210 cnh 1.11 \input{part1/atl6_figure}
211 cnh 1.3 %% CNHend
212    
213    
214 cnh 1.1 \subsection{Global ocean circulation}
215    
216 cnh 1.7 Figure \ref{fig:large-scale-circ} (top) shows the pattern of ocean currents at
217     the surface of a 4$^{\circ }$
218 cnh 1.2 global ocean model run with 15 vertical levels. Lopped cells are used to
219     represent topography on a regular $lat-lon$ grid extending from 70$^{\circ
220     }N $ to 70$^{\circ }S$. The model is driven using monthly-mean winds with
221     mixed boundary conditions on temperature and salinity at the surface. The
222     transfer properties of ocean eddies, convection and mixing is parameterized
223     in this model.
224    
225 cnh 1.7 Figure \ref{fig:large-scale-circ} (bottom) shows the meridional overturning
226     circulation of the global ocean in Sverdrups.
227 cnh 1.2
228 cnh 1.3 %%CNHbegin
229     \input{part1/global_circ_figure}
230     %%CNHend
231    
232 cnh 1.2 \subsection{Convection and mixing over topography}
233    
234     Dense plumes generated by localized cooling on the continental shelf of the
235     ocean may be influenced by rotation when the deformation radius is smaller
236     than the width of the cooling region. Rather than gravity plumes, the
237     mechanism for moving dense fluid down the shelf is then through geostrophic
238 adcroft 1.9 eddies. The simulation shown in the figure \ref{fig:convect-and-topo}
239 cnh 1.7 (blue is cold dense fluid, red is
240 cnh 1.2 warmer, lighter fluid) employs the non-hydrostatic capability of MITgcm to
241     trigger convection by surface cooling. The cold, dense water falls down the
242     slope but is deflected along the slope by rotation. It is found that
243     entrainment in the vertical plane is reduced when rotational control is
244     strong, and replaced by lateral entrainment due to the baroclinic
245     instability of the along-slope current.
246 cnh 1.1
247 cnh 1.3 %%CNHbegin
248     \input{part1/convect_and_topo}
249     %%CNHend
250    
251 cnh 1.1 \subsection{Boundary forced internal waves}
252    
253 cnh 1.2 The unique ability of MITgcm to treat non-hydrostatic dynamics in the
254     presence of complex geometry makes it an ideal tool to study internal wave
255     dynamics and mixing in oceanic canyons and ridges driven by large amplitude
256     barotropic tidal currents imposed through open boundary conditions.
257    
258 cnh 1.7 Fig. \ref{fig:boundary-forced-wave} shows the influence of cross-slope
259     topographic variations on
260 cnh 1.2 internal wave breaking - the cross-slope velocity is in color, the density
261     contoured. The internal waves are excited by application of open boundary
262 cnh 1.7 conditions on the left. They propagate to the sloping boundary (represented
263 cnh 1.2 using MITgcm's finite volume spatial discretization) where they break under
264     nonhydrostatic dynamics.
265    
266 cnh 1.3 %%CNHbegin
267     \input{part1/boundary_forced_waves}
268     %%CNHend
269    
270 cnh 1.2 \subsection{Parameter sensitivity using the adjoint of MITgcm}
271    
272     Forward and tangent linear counterparts of MITgcm are supported using an
273     `automatic adjoint compiler'. These can be used in parameter sensitivity and
274     data assimilation studies.
275    
276 cnh 1.7 As one example of application of the MITgcm adjoint, Figure \ref{fig:hf-sensitivity}
277     maps the gradient $\frac{\partial J}{\partial \mathcal{H}}$where $J$ is the magnitude
278 cnh 1.10 of the overturning stream-function shown in figure \ref{fig:large-scale-circ}
279 cnh 1.7 at 60$^{\circ }$N and $
280     \mathcal{H}(\lambda,\varphi)$ is the mean, local air-sea heat flux over
281     a 100 year period. We see that $J$ is
282 cnh 1.2 sensitive to heat fluxes over the Labrador Sea, one of the important sources
283     of deep water for the thermohaline circulations. This calculation also
284     yields sensitivities to all other model parameters.
285    
286 cnh 1.3 %%CNHbegin
287     \input{part1/adj_hf_ocean_figure}
288     %%CNHend
289    
290 cnh 1.2 \subsection{Global state estimation of the ocean}
291    
292     An important application of MITgcm is in state estimation of the global
293     ocean circulation. An appropriately defined `cost function', which measures
294     the departure of the model from observations (both remotely sensed and
295 cnh 1.10 in-situ) over an interval of time, is minimized by adjusting `control
296 cnh 1.2 parameters' such as air-sea fluxes, the wind field, the initial conditions
297 cnh 1.7 etc. Figure \ref{fig:assimilated-globes} shows an estimate of the time-mean
298     surface elevation of the ocean obtained by bringing the model in to
299     consistency with altimetric and in-situ observations over the period
300     1992-1997. {\bf CHANGE THIS TEXT - FIG FROM PATRICK/CARL/DETLEF}
301 cnh 1.2
302 cnh 1.3 %% CNHbegin
303 cnh 1.13 \input{part1/assim_figure}
304 cnh 1.3 %% CNHend
305    
306 cnh 1.2 \subsection{Ocean biogeochemical cycles}
307    
308     MITgcm is being used to study global biogeochemical cycles in the ocean. For
309     example one can study the effects of interannual changes in meteorological
310     forcing and upper ocean circulation on the fluxes of carbon dioxide and
311 cnh 1.7 oxygen between the ocean and atmosphere. Figure \ref{fig:biogeo} shows
312     the annual air-sea flux of oxygen and its relation to density outcrops in
313     the southern oceans from a single year of a global, interannually varying
314     simulation. The simulation is run at $1^{\circ}\times1^{\circ}$ resolution
315     telescoping to $\frac{1}{3}^{\circ}\times\frac{1}{3}^{\circ}$ in the tropics (not shown).
316 cnh 1.2
317 cnh 1.3 %%CNHbegin
318     \input{part1/biogeo_figure}
319     %%CNHend
320 cnh 1.2
321     \subsection{Simulations of laboratory experiments}
322    
323 cnh 1.7 Figure \ref{fig:lab-simulation} shows MITgcm being used to simulate a
324 cnh 1.10 laboratory experiment inquiring in to the dynamics of the Antarctic Circumpolar Current (ACC). An
325 cnh 1.2 initially homogeneous tank of water ($1m$ in diameter) is driven from its
326     free surface by a rotating heated disk. The combined action of mechanical
327     and thermal forcing creates a lens of fluid which becomes baroclinically
328     unstable. The stratification and depth of penetration of the lens is
329 cnh 1.7 arrested by its instability in a process analogous to that which sets the
330 cnh 1.2 stratification of the ACC.
331 cnh 1.1
332 cnh 1.3 %%CNHbegin
333     \input{part1/lab_figure}
334     %%CNHend
335    
336 cnh 1.13 % $Header: /u/u0/gcmpack/mitgcmdoc/part1/manual.tex,v 1.12 2001/11/19 14:33:44 cnh Exp $
337 cnh 1.1 % $Name: $
338    
339     \section{Continuous equations in `r' coordinates}
340    
341     To render atmosphere and ocean models from one dynamical core we exploit
342     `isomorphisms' between equation sets that govern the evolution of the
343 cnh 1.7 respective fluids - see figure \ref{fig:isomorphic-equations}.
344     One system of hydrodynamical equations is written down
345 cnh 1.1 and encoded. The model variables have different interpretations depending on
346     whether the atmosphere or ocean is being studied. Thus, for example, the
347     vertical coordinate `$r$' is interpreted as pressure, $p$, if we are
348 cnh 1.7 modeling the atmosphere (left hand side of figure \ref{fig:isomorphic-equations})
349     and height, $z$, if we are modeling the ocean (right hand side of figure
350     \ref{fig:isomorphic-equations}).
351 cnh 1.1
352 cnh 1.3 %%CNHbegin
353     \input{part1/zandpcoord_figure.tex}
354     %%CNHend
355    
356 cnh 1.1 The state of the fluid at any time is characterized by the distribution of
357     velocity $\vec{\mathbf{v}}$, active tracers $\theta $ and $S$, a
358     `geopotential' $\phi $ and density $\rho =\rho (\theta ,S,p)$ which may
359     depend on $\theta $, $S$, and $p$. The equations that govern the evolution
360     of these fields, obtained by applying the laws of classical mechanics and
361     thermodynamics to a Boussinesq, Navier-Stokes fluid are, written in terms of
362 cnh 1.7 a generic vertical coordinate, $r$, so that the appropriate
363     kinematic boundary conditions can be applied isomorphically
364     see figure \ref{fig:zandp-vert-coord}.
365 cnh 1.1
366 cnh 1.3 %%CNHbegin
367     \input{part1/vertcoord_figure.tex}
368     %%CNHend
369    
370 cnh 1.1 \begin{equation*}
371 adcroft 1.4 \frac{D\vec{\mathbf{v}_{h}}}{Dt}+\left( 2\vec{\Omega}\times \vec{\mathbf{v}}
372     \right) _{h}+\mathbf{\nabla }_{h}\phi =\mathcal{F}_{\vec{\mathbf{v}_{h}}}
373 cnh 1.8 \text{ horizontal mtm} \label{eq:horizontal_mtm}
374 cnh 1.1 \end{equation*}
375    
376 cnh 1.8 \begin{equation}
377 adcroft 1.4 \frac{D\dot{r}}{Dt}+\widehat{k}\cdot \left( 2\vec{\Omega}\times \vec{\mathbf{
378 cnh 1.1 v}}\right) +\frac{\partial \phi }{\partial r}+b=\mathcal{F}_{\dot{r}}\text{
379 cnh 1.8 vertical mtm} \label{eq:vertical_mtm}
380     \end{equation}
381 cnh 1.1
382     \begin{equation}
383 adcroft 1.4 \mathbf{\nabla }_{h}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \dot{r}}{
384 cnh 1.8 \partial r}=0\text{ continuity} \label{eq:continuity}
385 cnh 1.1 \end{equation}
386    
387 cnh 1.8 \begin{equation}
388     b=b(\theta ,S,r)\text{ equation of state} \label{eq:equation_of_state}
389     \end{equation}
390 cnh 1.1
391 cnh 1.8 \begin{equation}
392 cnh 1.2 \frac{D\theta }{Dt}=\mathcal{Q}_{\theta }\text{ potential temperature}
393 cnh 1.8 \label{eq:potential_temperature}
394     \end{equation}
395 cnh 1.1
396 cnh 1.8 \begin{equation}
397 cnh 1.2 \frac{DS}{Dt}=\mathcal{Q}_{S}\text{ humidity/salinity}
398 adcroft 1.9 \label{eq:humidity_salt}
399 cnh 1.8 \end{equation}
400 cnh 1.1
401     Here:
402    
403     \begin{equation*}
404 cnh 1.2 r\text{ is the vertical coordinate}
405 cnh 1.1 \end{equation*}
406    
407     \begin{equation*}
408     \frac{D}{Dt}=\frac{\partial }{\partial t}+\vec{\mathbf{v}}\cdot \nabla \text{
409 cnh 1.2 is the total derivative}
410 cnh 1.1 \end{equation*}
411    
412     \begin{equation*}
413 adcroft 1.4 \mathbf{\nabla }=\mathbf{\nabla }_{h}+\widehat{k}\frac{\partial }{\partial r}
414 cnh 1.2 \text{ is the `grad' operator}
415 cnh 1.1 \end{equation*}
416 adcroft 1.4 with $\mathbf{\nabla }_{h}$ operating in the horizontal and $\widehat{k}
417 cnh 1.1 \frac{\partial }{\partial r}$ operating in the vertical, where $\widehat{k}$
418     is a unit vector in the vertical
419    
420     \begin{equation*}
421 cnh 1.2 t\text{ is time}
422 cnh 1.1 \end{equation*}
423    
424     \begin{equation*}
425     \vec{\mathbf{v}}=(u,v,\dot{r})=(\vec{\mathbf{v}}_{h},\dot{r})\text{ is the
426 cnh 1.2 velocity}
427 cnh 1.1 \end{equation*}
428    
429     \begin{equation*}
430 cnh 1.2 \phi \text{ is the `pressure'/`geopotential'}
431 cnh 1.1 \end{equation*}
432    
433     \begin{equation*}
434 cnh 1.2 \vec{\Omega}\text{ is the Earth's rotation}
435 cnh 1.1 \end{equation*}
436    
437     \begin{equation*}
438 cnh 1.2 b\text{ is the `buoyancy'}
439 cnh 1.1 \end{equation*}
440    
441     \begin{equation*}
442 cnh 1.2 \theta \text{ is potential temperature}
443 cnh 1.1 \end{equation*}
444    
445     \begin{equation*}
446 cnh 1.2 S\text{ is specific humidity in the atmosphere; salinity in the ocean}
447 cnh 1.1 \end{equation*}
448    
449     \begin{equation*}
450 adcroft 1.4 \mathcal{F}_{\vec{\mathbf{v}}}\text{ are forcing and dissipation of }\vec{
451 cnh 1.1 \mathbf{v}}
452     \end{equation*}
453    
454     \begin{equation*}
455 cnh 1.2 \mathcal{Q}_{\theta }\mathcal{\ }\text{are forcing and dissipation of }\theta
456 cnh 1.1 \end{equation*}
457    
458     \begin{equation*}
459     \mathcal{Q}_{S}\mathcal{\ }\text{are forcing and dissipation of }S
460     \end{equation*}
461    
462     The $\mathcal{F}^{\prime }s$ and $\mathcal{Q}^{\prime }s$ are provided by
463 cnh 1.7 `physics' and forcing packages for atmosphere and ocean. These are described
464     in later chapters.
465 cnh 1.1
466     \subsection{Kinematic Boundary conditions}
467    
468     \subsubsection{vertical}
469    
470 cnh 1.7 at fixed and moving $r$ surfaces we set (see figure \ref{fig:zandp-vert-coord}):
471 cnh 1.1
472     \begin{equation}
473     \dot{r}=0atr=R_{fixed}(x,y)\text{ (ocean bottom, top of the atmosphere)}
474     \label{eq:fixedbc}
475     \end{equation}
476    
477     \begin{equation}
478     \dot{r}=\frac{Dr}{Dt}atr=R_{moving}\text{ \
479 cnh 1.10 (ocean surface,bottom of the atmosphere)} \label{eq:movingbc}
480 cnh 1.1 \end{equation}
481    
482     Here
483    
484     \begin{equation*}
485 cnh 1.2 R_{moving}=R_{o}+\eta
486 cnh 1.1 \end{equation*}
487     where $R_{o}(x,y)$ is the `$r-$value' (height or pressure, depending on
488     whether we are in the atmosphere or ocean) of the `moving surface' in the
489     resting fluid and $\eta $ is the departure from $R_{o}(x,y)$ in the presence
490     of motion.
491    
492     \subsubsection{horizontal}
493    
494     \begin{equation}
495     \vec{\mathbf{v}}\cdot \vec{\mathbf{n}}=0 \label{eq:noflow}
496 adcroft 1.4 \end{equation}
497 cnh 1.1 where $\vec{\mathbf{n}}$ is the normal to a solid boundary.
498    
499     \subsection{Atmosphere}
500    
501 cnh 1.7 In the atmosphere, (see figure \ref{fig:zandp-vert-coord}), we interpret:
502 cnh 1.1
503     \begin{equation}
504     r=p\text{ is the pressure} \label{eq:atmos-r}
505     \end{equation}
506    
507     \begin{equation}
508     \dot{r}=\frac{Dp}{Dt}=\omega \text{ is the vertical velocity in }p\text{
509     coordinates} \label{eq:atmos-omega}
510     \end{equation}
511    
512     \begin{equation}
513     \phi =g\,z\text{ is the geopotential height} \label{eq:atmos-phi}
514     \end{equation}
515    
516     \begin{equation}
517     b=\frac{\partial \Pi }{\partial p}\theta \text{ is the buoyancy}
518     \label{eq:atmos-b}
519     \end{equation}
520    
521     \begin{equation}
522     \theta =T(\frac{p_{c}}{p})^{\kappa }\text{ is potential temperature}
523     \label{eq:atmos-theta}
524     \end{equation}
525    
526     \begin{equation}
527     S=q,\text{ is the specific humidity} \label{eq:atmos-s}
528     \end{equation}
529     where
530    
531     \begin{equation*}
532     T\text{ is absolute temperature}
533 adcroft 1.4 \end{equation*}
534 cnh 1.1 \begin{equation*}
535     p\text{ is the pressure}
536 adcroft 1.4 \end{equation*}
537 cnh 1.1 \begin{eqnarray*}
538     &&z\text{ is the height of the pressure surface} \\
539     &&g\text{ is the acceleration due to gravity}
540     \end{eqnarray*}
541    
542     In the above the ideal gas law, $p=\rho RT$, has been expressed in terms of
543     the Exner function $\Pi (p)$ given by (see Appendix Atmosphere)
544     \begin{equation}
545     \Pi (p)=c_{p}(\frac{p}{p_{c}})^{\kappa } \label{eq:exner}
546 adcroft 1.4 \end{equation}
547 cnh 1.1 where $p_{c}$ is a reference pressure and $\kappa =R/c_{p}$ with $R$ the gas
548     constant and $c_{p}$ the specific heat of air at constant pressure.
549    
550     At the top of the atmosphere (which is `fixed' in our $r$ coordinate):
551    
552     \begin{equation*}
553 cnh 1.2 R_{fixed}=p_{top}=0
554 cnh 1.1 \end{equation*}
555     In a resting atmosphere the elevation of the mountains at the bottom is
556     given by
557     \begin{equation*}
558 cnh 1.2 R_{moving}=R_{o}(x,y)=p_{o}(x,y)
559 cnh 1.1 \end{equation*}
560     i.e. the (hydrostatic) pressure at the top of the mountains in a resting
561     atmosphere.
562    
563     The boundary conditions at top and bottom are given by:
564    
565     \begin{eqnarray}
566     &&\omega =0~\text{at }r=R_{fixed} \text{ (top of the atmosphere)}
567     \label{eq:fixed-bc-atmos} \\
568     \omega &=&\frac{Dp_{s}}{Dt}\text{; at }r=R_{moving}\text{ (bottom of the
569     atmosphere)} \label{eq:moving-bc-atmos}
570     \end{eqnarray}
571    
572 adcroft 1.9 Then the (hydrostatic form of) equations (\ref{eq:horizontal_mtm}-\ref{eq:humidity_salt})
573 cnh 1.8 yields a consistent set of atmospheric equations which, for convenience, are written out in $p$
574 cnh 1.1 coordinates in Appendix Atmosphere - see eqs(\ref{eq:atmos-prime}).
575    
576     \subsection{Ocean}
577    
578     In the ocean we interpret:
579     \begin{eqnarray}
580     r &=&z\text{ is the height} \label{eq:ocean-z} \\
581     \dot{r} &=&\frac{Dz}{Dt}=w\text{ is the vertical velocity}
582     \label{eq:ocean-w} \\
583     \phi &=&\frac{p}{\rho _{c}}\text{ is the pressure} \label{eq:ocean-p} \\
584     b(\theta ,S,r) &=&\frac{g}{\rho _{c}}\left( \rho (\theta ,S,r)-\rho
585     _{c}\right) \text{ is the buoyancy} \label{eq:ocean-b}
586     \end{eqnarray}
587     where $\rho _{c}$ is a fixed reference density of water and $g$ is the
588     acceleration due to gravity.\noindent
589    
590     In the above
591    
592     At the bottom of the ocean: $R_{fixed}(x,y)=-H(x,y)$.
593    
594     The surface of the ocean is given by: $R_{moving}=\eta $
595    
596 adcroft 1.4 The position of the resting free surface of the ocean is given by $
597 cnh 1.1 R_{o}=Z_{o}=0$.
598    
599     Boundary conditions are:
600    
601     \begin{eqnarray}
602     w &=&0~\text{at }r=R_{fixed}\text{ (ocean bottom)} \label{eq:fixed-bc-ocean}
603     \\
604 adcroft 1.4 w &=&\frac{D\eta }{Dt}\text{ at }r=R_{moving}=\eta \text{ (ocean surface)
605 cnh 1.1 \label{eq:moving-bc-ocean}}
606     \end{eqnarray}
607     where $\eta $ is the elevation of the free surface.
608    
609 adcroft 1.9 Then equations (\ref{eq:horizontal_mtm}-\ref{eq:humidity_salt}) yield a consistent set
610 cnh 1.8 of oceanic equations
611 cnh 1.1 which, for convenience, are written out in $z$ coordinates in Appendix Ocean
612     - see eqs(\ref{eq:ocean-mom}) to (\ref{eq:ocean-salt}).
613    
614     \subsection{Hydrostatic, Quasi-hydrostatic, Quasi-nonhydrostatic and
615     Non-hydrostatic forms}
616    
617     Let us separate $\phi $ in to surface, hydrostatic and non-hydrostatic terms:
618    
619     \begin{equation}
620     \phi (x,y,r)=\phi _{s}(x,y)+\phi _{hyd}(x,y,r)+\phi _{nh}(x,y,r)
621     \label{eq:phi-split}
622 adcroft 1.4 \end{equation}
623 cnh 1.8 and write eq(\ref{eq:incompressible}) in the form:
624 cnh 1.1
625     \begin{equation}
626     \frac{\partial \vec{\mathbf{v}_{h}}}{\partial t}+\mathbf{\nabla }_{h}\phi
627     _{s}+\mathbf{\nabla }_{h}\phi _{hyd}+\epsilon _{nh}\mathbf{\nabla }_{h}\phi
628     _{nh}=\vec{\mathbf{G}}_{\vec{v}_{h}} \label{eq:mom-h}
629     \end{equation}
630    
631     \begin{equation}
632     \frac{\partial \phi _{hyd}}{\partial r}=-b \label{eq:hydrostatic}
633     \end{equation}
634    
635     \begin{equation}
636 adcroft 1.4 \epsilon _{nh}\frac{\partial \dot{r}}{\partial t}+\frac{\partial \phi _{nh}}{
637 cnh 1.1 \partial r}=G_{\dot{r}} \label{eq:mom-w}
638     \end{equation}
639     Here $\epsilon _{nh}$ is a non-hydrostatic parameter.
640    
641 adcroft 1.4 The $\left( \vec{\mathbf{G}}_{\vec{v}},G_{\dot{r}}\right) $ in eq(\ref
642 cnh 1.1 {eq:mom-h}) and (\ref{eq:mom-w}) represent advective, metric and Coriolis
643 adcroft 1.4 terms in the momentum equations. In spherical coordinates they take the form
644     \footnote{
645 cnh 1.1 In the hydrostatic primitive equations (\textbf{HPE}) all underlined terms
646 adcroft 1.4 in (\ref{eq:gu-speherical}), (\ref{eq:gv-spherical}) and (\ref
647 cnh 1.1 {eq:gw-spherical}) are omitted; the singly-underlined terms are included in
648 adcroft 1.4 the quasi-hydrostatic model (\textbf{QH}). The fully non-hydrostatic model (
649 cnh 1.1 \textbf{NH}) includes all terms.} - see Marshall et al 1997a for a full
650     discussion:
651    
652     \begin{equation}
653     \left.
654     \begin{tabular}{l}
655     $G_{u}=-\vec{\mathbf{v}}.\nabla u$ \\
656 cnh 1.6 $-\left\{ \underline{\frac{u\dot{r}}{{r}}}-\frac{uv\tan \varphi}{{r}}\right\} $
657 cnh 1.1 \\
658 cnh 1.6 $-\left\{ -2\Omega v\sin \varphi+\underline{2\Omega \dot{r}\cos \varphi}\right\} $
659 cnh 1.1 \\
660 adcroft 1.4 $+\mathcal{F}_{u}$
661     \end{tabular}
662 cnh 1.1 \ \right\} \left\{
663     \begin{tabular}{l}
664     \textit{advection} \\
665     \textit{metric} \\
666     \textit{Coriolis} \\
667 adcroft 1.4 \textit{\ Forcing/Dissipation}
668     \end{tabular}
669 cnh 1.2 \ \right. \qquad \label{eq:gu-speherical}
670 cnh 1.1 \end{equation}
671    
672     \begin{equation}
673     \left.
674     \begin{tabular}{l}
675     $G_{v}=-\vec{\mathbf{v}}.\nabla v$ \\
676 cnh 1.6 $-\left\{ \underline{\frac{v\dot{r}}{{r}}}-\frac{u^{2}\tan \varphi}{{r}}\right\}
677 cnh 1.1 $ \\
678 cnh 1.6 $-\left\{ -2\Omega u\sin \varphi \right\} $ \\
679 adcroft 1.4 $+\mathcal{F}_{v}$
680     \end{tabular}
681 cnh 1.1 \ \right\} \left\{
682     \begin{tabular}{l}
683     \textit{advection} \\
684     \textit{metric} \\
685     \textit{Coriolis} \\
686 adcroft 1.4 \textit{\ Forcing/Dissipation}
687     \end{tabular}
688 cnh 1.2 \ \right. \qquad \label{eq:gv-spherical}
689 adcroft 1.4 \end{equation}
690 cnh 1.2 \qquad \qquad \qquad \qquad \qquad
691 cnh 1.1
692     \begin{equation}
693     \left.
694     \begin{tabular}{l}
695     $G_{\dot{r}}=-\underline{\underline{\vec{\mathbf{v}}.\nabla \dot{r}}}$ \\
696     $+\left\{ \underline{\frac{u^{_{^{2}}}+v^{2}}{{r}}}\right\} $ \\
697 cnh 1.6 ${+}\underline{{2\Omega u\cos \varphi}}$ \\
698 adcroft 1.4 $\underline{\underline{\mathcal{F}_{\dot{r}}}}$
699     \end{tabular}
700 cnh 1.1 \ \right\} \left\{
701     \begin{tabular}{l}
702     \textit{advection} \\
703     \textit{metric} \\
704     \textit{Coriolis} \\
705 adcroft 1.4 \textit{\ Forcing/Dissipation}
706     \end{tabular}
707 cnh 1.2 \ \right. \label{eq:gw-spherical}
708 adcroft 1.4 \end{equation}
709 cnh 1.2 \qquad \qquad \qquad \qquad \qquad
710 cnh 1.1
711 cnh 1.6 In the above `${r}$' is the distance from the center of the earth and `$\varphi$
712 cnh 1.1 ' is latitude.
713    
714     Grad and div operators in spherical coordinates are defined in appendix
715 adcroft 1.4 OPERATORS.
716 cnh 1.1
717 cnh 1.3 %%CNHbegin
718     \input{part1/sphere_coord_figure.tex}
719     %%CNHend
720    
721 cnh 1.1 \subsubsection{Shallow atmosphere approximation}
722    
723     Most models are based on the `hydrostatic primitive equations' (HPE's) in
724     which the vertical momentum equation is reduced to a statement of
725     hydrostatic balance and the `traditional approximation' is made in which the
726     Coriolis force is treated approximately and the shallow atmosphere
727     approximation is made.\ The MITgcm need not make the `traditional
728     approximation'. To be able to support consistent non-hydrostatic forms the
729 adcroft 1.4 shallow atmosphere approximation can be relaxed - when dividing through by $
730 cnh 1.2 r $ in, for example, (\ref{eq:gu-speherical}), we do not replace $r$ by $a$,
731 cnh 1.1 the radius of the earth.
732    
733     \subsubsection{Hydrostatic and quasi-hydrostatic forms}
734 cnh 1.7 \label{sec:hydrostatic_and_quasi-hydrostatic_forms}
735 cnh 1.1
736     These are discussed at length in Marshall et al (1997a).
737    
738     In the `hydrostatic primitive equations' (\textbf{HPE)} all the underlined
739     terms in Eqs. (\ref{eq:gu-speherical} $\rightarrow $\ \ref{eq:gw-spherical})
740     are neglected and `${r}$' is replaced by `$a$', the mean radius of the
741     earth. Once the pressure is found at one level - e.g. by inverting a 2-d
742     Elliptic equation for $\phi _{s}$ at $r=R_{moving}$ - the pressure can be
743 adcroft 1.4 computed at all other levels by integration of the hydrostatic relation, eq(
744 cnh 1.1 \ref{eq:hydrostatic}).
745    
746     In the `quasi-hydrostatic' equations (\textbf{QH)} strict balance between
747     gravity and vertical pressure gradients is not imposed. The $2\Omega u\cos
748 cnh 1.6 \varphi $ Coriolis term are not neglected and are balanced by a non-hydrostatic
749 adcroft 1.4 contribution to the pressure field: only the terms underlined twice in Eqs. (
750 cnh 1.1 \ref{eq:gu-speherical}$\rightarrow $\ \ref{eq:gw-spherical}) are set to zero
751     and, simultaneously, the shallow atmosphere approximation is relaxed. In
752     \textbf{QH}\ \textit{all} the metric terms are retained and the full
753     variation of the radial position of a particle monitored. The \textbf{QH}\
754     vertical momentum equation (\ref{eq:mom-w}) becomes:
755    
756     \begin{equation*}
757 cnh 1.6 \frac{\partial \phi _{nh}}{\partial r}=2\Omega u\cos \varphi
758 cnh 1.1 \end{equation*}
759     making a small correction to the hydrostatic pressure.
760    
761     \textbf{QH} has good energetic credentials - they are the same as for
762     \textbf{HPE}. Importantly, however, it has the same angular momentum
763     principle as the full non-hydrostatic model (\textbf{NH)} - see Marshall
764     et.al., 1997a. As in \textbf{HPE }only a 2-d elliptic problem need be solved.
765    
766     \subsubsection{Non-hydrostatic and quasi-nonhydrostatic forms}
767    
768     The MIT model presently supports a full non-hydrostatic ocean isomorph, but
769     only a quasi-non-hydrostatic atmospheric isomorph.
770    
771     \paragraph{Non-hydrostatic Ocean}
772    
773 adcroft 1.4 In the non-hydrostatic ocean model all terms in equations Eqs.(\ref
774 cnh 1.1 {eq:gu-speherical} $\rightarrow $\ \ref{eq:gw-spherical}) are retained. A
775     three dimensional elliptic equation must be solved subject to Neumann
776     boundary conditions (see below). It is important to note that use of the
777     full \textbf{NH} does not admit any new `fast' waves in to the system - the
778 cnh 1.8 incompressible condition eq(\ref{eq:continuity}) has already filtered out
779 cnh 1.1 acoustic modes. It does, however, ensure that the gravity waves are treated
780     accurately with an exact dispersion relation. The \textbf{NH} set has a
781     complete angular momentum principle and consistent energetics - see White
782     and Bromley, 1995; Marshall et.al.\ 1997a.
783    
784     \paragraph{Quasi-nonhydrostatic Atmosphere}
785    
786 adcroft 1.4 In the non-hydrostatic version of our atmospheric model we approximate $\dot{
787 cnh 1.1 r}$ in the vertical momentum eqs(\ref{eq:mom-w}) and (\ref{eq:gv-spherical})
788     (but only here) by:
789    
790     \begin{equation}
791     \dot{r}=\frac{Dp}{Dt}=\frac{1}{g}\frac{D\phi }{Dt} \label{eq:quasi-nh-w}
792 adcroft 1.4 \end{equation}
793 cnh 1.1 where $p_{hy}$ is the hydrostatic pressure.
794    
795     \subsubsection{Summary of equation sets supported by model}
796    
797     \paragraph{Atmosphere}
798    
799     Hydrostatic, and quasi-hydrostatic and quasi non-hydrostatic forms of the
800     compressible non-Boussinesq equations in $p-$coordinates are supported.
801    
802     \subparagraph{Hydrostatic and quasi-hydrostatic}
803    
804     The hydrostatic set is written out in $p-$coordinates in appendix Atmosphere
805     - see eq(\ref{eq:atmos-prime}).
806    
807     \subparagraph{Quasi-nonhydrostatic}
808    
809     A quasi-nonhydrostatic form is also supported.
810    
811     \paragraph{Ocean}
812    
813     \subparagraph{Hydrostatic and quasi-hydrostatic}
814    
815     Hydrostatic, and quasi-hydrostatic forms of the incompressible Boussinesq
816     equations in $z-$coordinates are supported.
817    
818     \subparagraph{Non-hydrostatic}
819    
820 adcroft 1.4 Non-hydrostatic forms of the incompressible Boussinesq equations in $z-$
821     coordinates are supported - see eqs(\ref{eq:ocean-mom}) to (\ref
822 cnh 1.1 {eq:ocean-salt}).
823    
824     \subsection{Solution strategy}
825    
826 adcroft 1.4 The method of solution employed in the \textbf{HPE}, \textbf{QH} and \textbf{
827 cnh 1.8 NH} models is summarized in Figure \ref{fig:solution-strategy}.
828     Under all dynamics, a 2-d elliptic equation is
829 cnh 1.1 first solved to find the surface pressure and the hydrostatic pressure at
830     any level computed from the weight of fluid above. Under \textbf{HPE} and
831     \textbf{QH} dynamics, the horizontal momentum equations are then stepped
832     forward and $\dot{r}$ found from continuity. Under \textbf{NH} dynamics a
833     3-d elliptic equation must be solved for the non-hydrostatic pressure before
834     stepping forward the horizontal momentum equations; $\dot{r}$ is found by
835     stepping forward the vertical momentum equation.
836    
837 cnh 1.3 %%CNHbegin
838     \input{part1/solution_strategy_figure.tex}
839     %%CNHend
840    
841 cnh 1.1 There is no penalty in implementing \textbf{QH} over \textbf{HPE} except, of
842 cnh 1.6 course, some complication that goes with the inclusion of $\cos \varphi \ $
843 cnh 1.1 Coriolis terms and the relaxation of the shallow atmosphere approximation.
844     But this leads to negligible increase in computation. In \textbf{NH}, in
845     contrast, one additional elliptic equation - a three-dimensional one - must
846     be inverted for $p_{nh}$. However the `overhead' of the \textbf{NH} model is
847     essentially negligible in the hydrostatic limit (see detailed discussion in
848     Marshall et al, 1997) resulting in a non-hydrostatic algorithm that, in the
849     hydrostatic limit, is as computationally economic as the \textbf{HPEs}.
850    
851     \subsection{Finding the pressure field}
852 cnh 1.7 \label{sec:finding_the_pressure_field}
853 cnh 1.1
854     Unlike the prognostic variables $u$, $v$, $w$, $\theta $ and $S$, the
855     pressure field must be obtained diagnostically. We proceed, as before, by
856     dividing the total (pressure/geo) potential in to three parts, a surface
857     part, $\phi _{s}(x,y)$, a hydrostatic part $\phi _{hyd}(x,y,r)$ and a
858     non-hydrostatic part $\phi _{nh}(x,y,r)$, as in (\ref{eq:phi-split}), and
859     writing the momentum equation as in (\ref{eq:mom-h}).
860    
861     \subsubsection{Hydrostatic pressure}
862    
863     Hydrostatic pressure is obtained by integrating (\ref{eq:hydrostatic})
864     vertically from $r=R_{o}$ where $\phi _{hyd}(r=R_{o})=0$, to yield:
865    
866     \begin{equation*}
867 adcroft 1.4 \int_{r}^{R_{o}}\frac{\partial \phi _{hyd}}{\partial r}dr=\left[ \phi _{hyd}
868 cnh 1.2 \right] _{r}^{R_{o}}=\int_{r}^{R_{o}}-bdr
869 cnh 1.1 \end{equation*}
870     and so
871    
872     \begin{equation}
873     \phi _{hyd}(x,y,r)=\int_{r}^{R_{o}}bdr \label{eq:hydro-phi}
874     \end{equation}
875    
876     The model can be easily modified to accommodate a loading term (e.g
877     atmospheric pressure pushing down on the ocean's surface) by setting:
878    
879     \begin{equation}
880     \phi _{hyd}(r=R_{o})=loading \label{eq:loading}
881     \end{equation}
882    
883     \subsubsection{Surface pressure}
884    
885 cnh 1.8 The surface pressure equation can be obtained by integrating continuity,
886     (\ref{eq:continuity}), vertically from $r=R_{fixed}$ to $r=R_{moving}$
887 cnh 1.1
888     \begin{equation*}
889 adcroft 1.4 \int_{R_{fixed}}^{R_{moving}}\left( \mathbf{\nabla }_{h}\cdot \vec{\mathbf{v}
890 cnh 1.2 }_{h}+\partial _{r}\dot{r}\right) dr=0
891 cnh 1.1 \end{equation*}
892    
893     Thus:
894    
895     \begin{equation*}
896     \frac{\partial \eta }{\partial t}+\vec{\mathbf{v}}.\nabla \eta
897 adcroft 1.4 +\int_{R_{fixed}}^{R_{moving}}\mathbf{\nabla }_{h}\cdot \vec{\mathbf{v}}
898 cnh 1.2 _{h}dr=0
899 cnh 1.1 \end{equation*}
900 adcroft 1.4 where $\eta =R_{moving}-R_{o}$ is the free-surface $r$-anomaly in units of $
901 cnh 1.1 r $. The above can be rearranged to yield, using Leibnitz's theorem:
902    
903     \begin{equation}
904     \frac{\partial \eta }{\partial t}+\mathbf{\nabla }_{h}\cdot
905     \int_{R_{fixed}}^{R_{moving}}\vec{\mathbf{v}}_{h}dr=\text{source}
906     \label{eq:free-surface}
907 adcroft 1.4 \end{equation}
908 cnh 1.1 where we have incorporated a source term.
909    
910     Whether $\phi $ is pressure (ocean model, $p/\rho _{c}$) or geopotential
911 cnh 1.8 (atmospheric model), in (\ref{eq:mom-h}), the horizontal gradient term can
912 cnh 1.1 be written
913     \begin{equation}
914 cnh 1.2 \mathbf{\nabla }_{h}\phi _{s}=\mathbf{\nabla }_{h}\left( b_{s}\eta \right)
915 cnh 1.1 \label{eq:phi-surf}
916 adcroft 1.4 \end{equation}
917 cnh 1.1 where $b_{s}$ is the buoyancy at the surface.
918    
919 cnh 1.8 In the hydrostatic limit ($\epsilon _{nh}=0$), equations (\ref{eq:mom-h}), (\ref
920 cnh 1.1 {eq:free-surface}) and (\ref{eq:phi-surf}) can be solved by inverting a 2-d
921     elliptic equation for $\phi _{s}$ as described in Chapter 2. Both `free
922     surface' and `rigid lid' approaches are available.
923    
924     \subsubsection{Non-hydrostatic pressure}
925    
926 cnh 1.8 Taking the horizontal divergence of (\ref{eq:mom-h}) and adding
927     $\frac{\partial }{\partial r}$ of (\ref{eq:mom-w}), invoking the continuity equation
928     (\ref{eq:continuity}), we deduce that:
929 cnh 1.1
930     \begin{equation}
931 adcroft 1.4 \nabla _{3}^{2}\phi _{nh}=\nabla .\vec{\mathbf{G}}_{\vec{v}}-\left( \mathbf{
932     \nabla }_{h}^{2}\phi _{s}+\mathbf{\nabla }^{2}\phi _{hyd}\right) =\nabla .
933 cnh 1.1 \vec{\mathbf{F}} \label{eq:3d-invert}
934     \end{equation}
935    
936     For a given rhs this 3-d elliptic equation must be inverted for $\phi _{nh}$
937     subject to appropriate choice of boundary conditions. This method is usually
938     called \textit{The Pressure Method} [Harlow and Welch, 1965; Williams, 1969;
939     Potter, 1976]. In the hydrostatic primitive equations case (\textbf{HPE}),
940     the 3-d problem does not need to be solved.
941    
942     \paragraph{Boundary Conditions}
943    
944     We apply the condition of no normal flow through all solid boundaries - the
945     coasts (in the ocean) and the bottom:
946    
947     \begin{equation}
948     \vec{\mathbf{v}}.\widehat{n}=0 \label{nonormalflow}
949     \end{equation}
950     where $\widehat{n}$ is a vector of unit length normal to the boundary. The
951     kinematic condition (\ref{nonormalflow}) is also applied to the vertical
952 adcroft 1.4 velocity at $r=R_{moving}$. No-slip $\left( v_{T}=0\right) \ $or slip $
953 cnh 1.1 \left( \partial v_{T}/\partial n=0\right) \ $conditions are employed on the
954     tangential component of velocity, $v_{T}$, at all solid boundaries,
955     depending on the form chosen for the dissipative terms in the momentum
956     equations - see below.
957    
958 cnh 1.8 Eq.(\ref{nonormalflow}) implies, making use of (\ref{eq:mom-h}), that:
959 cnh 1.1
960     \begin{equation}
961     \widehat{n}.\nabla \phi _{nh}=\widehat{n}.\vec{\mathbf{F}}
962     \label{eq:inhom-neumann-nh}
963     \end{equation}
964     where
965    
966     \begin{equation*}
967     \vec{\mathbf{F}}=\vec{\mathbf{G}}_{\vec{v}}-\left( \mathbf{\nabla }_{h}\phi
968     _{s}+\mathbf{\nabla }\phi _{hyd}\right)
969 adcroft 1.4 \end{equation*}
970 cnh 1.1 presenting inhomogeneous Neumann boundary conditions to the Elliptic problem
971     (\ref{eq:3d-invert}). As shown, for example, by Williams (1969), one can
972     exploit classical 3D potential theory and, by introducing an appropriately
973 cnh 1.2 chosen $\delta $-function sheet of `source-charge', replace the
974     inhomogeneous boundary condition on pressure by a homogeneous one. The
975 adcroft 1.4 source term $rhs$ in (\ref{eq:3d-invert}) is the divergence of the vector $
976     \vec{\mathbf{F}}.$ By simultaneously setting $
977 cnh 1.1 \begin{array}{l}
978 adcroft 1.4 \widehat{n}.\vec{\mathbf{F}}
979     \end{array}
980 cnh 1.1 =0$\ and $\widehat{n}.\nabla \phi _{nh}=0\ $on the boundary the following
981 cnh 1.2 self-consistent but simpler homogenized Elliptic problem is obtained:
982 cnh 1.1
983     \begin{equation*}
984 cnh 1.2 \nabla ^{2}\phi _{nh}=\nabla .\widetilde{\vec{\mathbf{F}}}\qquad
985 adcroft 1.4 \end{equation*}
986 cnh 1.1 where $\widetilde{\vec{\mathbf{F}}}$ is a modified $\vec{\mathbf{F}}$ such
987 adcroft 1.4 that $\widetilde{\vec{\mathbf{F}}}.\widehat{n}=0$. As is implied by (\ref
988 cnh 1.1 {eq:inhom-neumann-nh}) the modified boundary condition becomes:
989    
990     \begin{equation}
991     \widehat{n}.\nabla \phi _{nh}=0 \label{eq:hom-neumann-nh}
992     \end{equation}
993    
994     If the flow is `close' to hydrostatic balance then the 3-d inversion
995     converges rapidly because $\phi _{nh}\ $is then only a small correction to
996     the hydrostatic pressure field (see the discussion in Marshall et al, a,b).
997    
998 cnh 1.8 The solution $\phi _{nh}\ $to (\ref{eq:3d-invert}) and (\ref{eq:inhom-neumann-nh})
999 cnh 1.1 does not vanish at $r=R_{moving}$, and so refines the pressure there.
1000    
1001     \subsection{Forcing/dissipation}
1002    
1003     \subsubsection{Forcing}
1004    
1005     The forcing terms $\mathcal{F}$ on the rhs of the equations are provided by
1006 cnh 1.8 `physics packages' and forcing packages. These are described later on.
1007 cnh 1.1
1008     \subsubsection{Dissipation}
1009    
1010     \paragraph{Momentum}
1011    
1012     Many forms of momentum dissipation are available in the model. Laplacian and
1013     biharmonic frictions are commonly used:
1014    
1015     \begin{equation}
1016 adcroft 1.4 D_{V}=A_{h}\nabla _{h}^{2}v+A_{v}\frac{\partial ^{2}v}{\partial z^{2}}
1017 cnh 1.1 +A_{4}\nabla _{h}^{4}v \label{eq:dissipation}
1018     \end{equation}
1019     where $A_{h}$ and $A_{v}\ $are (constant) horizontal and vertical viscosity
1020     coefficients and $A_{4}\ $is the horizontal coefficient for biharmonic
1021     friction. These coefficients are the same for all velocity components.
1022    
1023     \paragraph{Tracers}
1024    
1025     The mixing terms for the temperature and salinity equations have a similar
1026     form to that of momentum except that the diffusion tensor can be
1027 adcroft 1.4 non-diagonal and have varying coefficients. $\qquad $
1028 cnh 1.1 \begin{equation}
1029     D_{T,S}=\nabla .[\underline{\underline{K}}\nabla (T,S)]+K_{4}\nabla
1030     _{h}^{4}(T,S) \label{eq:diffusion}
1031     \end{equation}
1032 adcroft 1.4 where $\underline{\underline{K}}\ $is the diffusion tensor and the $K_{4}\ $
1033 cnh 1.1 horizontal coefficient for biharmonic diffusion. In the simplest case where
1034     the subgrid-scale fluxes of heat and salt are parameterized with constant
1035     horizontal and vertical diffusion coefficients, $\underline{\underline{K}}$,
1036     reduces to a diagonal matrix with constant coefficients:
1037    
1038     \begin{equation}
1039     \qquad \qquad \qquad \qquad K=\left(
1040     \begin{array}{ccc}
1041     K_{h} & 0 & 0 \\
1042     0 & K_{h} & 0 \\
1043 adcroft 1.4 0 & 0 & K_{v}
1044 cnh 1.1 \end{array}
1045     \right) \qquad \qquad \qquad \label{eq:diagonal-diffusion-tensor}
1046     \end{equation}
1047     where $K_{h}\ $and $K_{v}\ $are the horizontal and vertical diffusion
1048     coefficients. These coefficients are the same for all tracers (temperature,
1049     salinity ... ).
1050    
1051     \subsection{Vector invariant form}
1052    
1053 adcroft 1.4 For some purposes it is advantageous to write momentum advection in eq(\ref
1054 cnh 1.8 {eq:horizontal_mtm}) and (\ref{eq:vertical_mtm}) in the (so-called) `vector invariant' form:
1055 cnh 1.1
1056     \begin{equation}
1057 adcroft 1.4 \frac{D\vec{\mathbf{v}}}{Dt}=\frac{\partial \vec{\mathbf{v}}}{\partial t}
1058     +\left( \nabla \times \vec{\mathbf{v}}\right) \times \vec{\mathbf{v}}+\nabla
1059 cnh 1.2 \left[ \frac{1}{2}(\vec{\mathbf{v}}\cdot \vec{\mathbf{v}})\right]
1060 cnh 1.1 \label{eq:vi-identity}
1061 adcroft 1.4 \end{equation}
1062 cnh 1.1 This permits alternative numerical treatments of the non-linear terms based
1063     on their representation as a vorticity flux. Because gradients of coordinate
1064     vectors no longer appear on the rhs of (\ref{eq:vi-identity}), explicit
1065 adcroft 1.4 representation of the metric terms in (\ref{eq:gu-speherical}), (\ref
1066 cnh 1.1 {eq:gv-spherical}) and (\ref{eq:gw-spherical}), can be avoided: information
1067     about the geometry is contained in the areas and lengths of the volumes used
1068     to discretize the model.
1069    
1070     \subsection{Adjoint}
1071    
1072 cnh 1.8 Tangent linear and adjoint counterparts of the forward model are described
1073 cnh 1.2 in Chapter 5.
1074 cnh 1.1
1075 cnh 1.13 % $Header: /u/u0/gcmpack/mitgcmdoc/part1/manual.tex,v 1.12 2001/11/19 14:33:44 cnh Exp $
1076 cnh 1.1 % $Name: $
1077    
1078     \section{Appendix ATMOSPHERE}
1079    
1080     \subsection{Hydrostatic Primitive Equations for the Atmosphere in pressure
1081     coordinates}
1082    
1083     \label{sect-hpe-p}
1084    
1085     The hydrostatic primitive equations (HPEs) in p-coordinates are:
1086     \begin{eqnarray}
1087 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1088 cnh 1.2 _{h}+\mathbf{\nabla }_{p}\phi &=&\vec{\mathbf{\mathcal{F}}}
1089 cnh 1.1 \label{eq:atmos-mom} \\
1090 cnh 1.2 \frac{\partial \phi }{\partial p}+\alpha &=&0 \label{eq-p-hydro-start} \\
1091 adcroft 1.4 \mathbf{\nabla }_{p}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \omega }{
1092 cnh 1.1 \partial p} &=&0 \label{eq:atmos-cont} \\
1093 cnh 1.2 p\alpha &=&RT \label{eq:atmos-eos} \\
1094 cnh 1.1 c_{v}\frac{DT}{Dt}+p\frac{D\alpha }{Dt} &=&\mathcal{Q} \label{eq:atmos-heat}
1095 adcroft 1.4 \end{eqnarray}
1096 cnh 1.1 where $\vec{\mathbf{v}}_{h}=(u,v,0)$ is the `horizontal' (on pressure
1097     surfaces) component of velocity,$\frac{D}{Dt}=\vec{\mathbf{v}}_{h}\cdot
1098     \mathbf{\nabla }_{p}+\omega \frac{\partial }{\partial p}$ is the total
1099 cnh 1.6 derivative, $f=2\Omega \sin \varphi$ is the Coriolis parameter, $\phi =gz$ is
1100 adcroft 1.4 the geopotential, $\alpha =1/\rho $ is the specific volume, $\omega =\frac{Dp
1101     }{Dt}$ is the vertical velocity in the $p-$coordinate. Equation(\ref
1102     {eq:atmos-heat}) is the first law of thermodynamics where internal energy $
1103     e=c_{v}T$, $T$ is temperature, $Q$ is the rate of heating per unit mass and $
1104 cnh 1.1 p\frac{D\alpha }{Dt}$ is the work done by the fluid in compressing.
1105    
1106     It is convenient to cast the heat equation in terms of potential temperature
1107     $\theta $ so that it looks more like a generic conservation law.
1108     Differentiating (\ref{eq:atmos-eos}) we get:
1109     \begin{equation*}
1110     p\frac{D\alpha }{Dt}+\alpha \frac{Dp}{Dt}=R\frac{DT}{Dt}
1111 adcroft 1.4 \end{equation*}
1112     which, when added to the heat equation (\ref{eq:atmos-heat}) and using $
1113 cnh 1.1 c_{p}=c_{v}+R$, gives:
1114     \begin{equation}
1115     c_{p}\frac{DT}{Dt}-\alpha \frac{Dp}{Dt}=\mathcal{Q}
1116     \label{eq-p-heat-interim}
1117 adcroft 1.4 \end{equation}
1118 cnh 1.1 Potential temperature is defined:
1119     \begin{equation}
1120     \theta =T(\frac{p_{c}}{p})^{\kappa } \label{eq:potential-temp}
1121 adcroft 1.4 \end{equation}
1122 cnh 1.1 where $p_{c}$ is a reference pressure and $\kappa =R/c_{p}$. For convenience
1123     we will make use of the Exner function $\Pi (p)$ which defined by:
1124     \begin{equation}
1125     \Pi (p)=c_{p}(\frac{p}{p_{c}})^{\kappa } \label{Exner}
1126 adcroft 1.4 \end{equation}
1127 cnh 1.1 The following relations will be useful and are easily expressed in terms of
1128     the Exner function:
1129     \begin{equation*}
1130     c_{p}T=\Pi \theta \;\;;\;\;\frac{\partial \Pi }{\partial p}=\frac{\kappa \Pi
1131 adcroft 1.4 }{p}\;\;;\;\;\alpha =\frac{\kappa \Pi \theta }{p}=\frac{\partial \ \Pi }{
1132     \partial p}\theta \;\;;\;\;\frac{D\Pi }{Dt}=\frac{\partial \Pi }{\partial p}
1133 cnh 1.1 \frac{Dp}{Dt}
1134 adcroft 1.4 \end{equation*}
1135 cnh 1.1 where $b=\frac{\partial \ \Pi }{\partial p}\theta $ is the buoyancy.
1136    
1137     The heat equation is obtained by noting that
1138     \begin{equation*}
1139     c_{p}\frac{DT}{Dt}=\frac{D(\Pi \theta )}{Dt}=\Pi \frac{D\theta }{Dt}+\theta
1140 cnh 1.2 \frac{D\Pi }{Dt}=\Pi \frac{D\theta }{Dt}+\alpha \frac{Dp}{Dt}
1141 cnh 1.1 \end{equation*}
1142     and on substituting into (\ref{eq-p-heat-interim}) gives:
1143     \begin{equation}
1144     \Pi \frac{D\theta }{Dt}=\mathcal{Q}
1145     \label{eq:potential-temperature-equation}
1146     \end{equation}
1147     which is in conservative form.
1148    
1149 adcroft 1.4 For convenience in the model we prefer to step forward (\ref
1150 cnh 1.1 {eq:potential-temperature-equation}) rather than (\ref{eq:atmos-heat}).
1151    
1152     \subsubsection{Boundary conditions}
1153    
1154     The upper and lower boundary conditions are :
1155     \begin{eqnarray}
1156     \mbox{at the top:}\;\;p=0 &&\text{, }\omega =\frac{Dp}{Dt}=0 \\
1157     \mbox{at the surface:}\;\;p=p_{s} &&\text{, }\phi =\phi _{topo}=g~Z_{topo}
1158     \label{eq:boundary-condition-atmosphere}
1159     \end{eqnarray}
1160     In $p$-coordinates, the upper boundary acts like a solid boundary ($\omega
1161     =0 $); in $z$-coordinates and the lower boundary is analogous to a free
1162     surface ($\phi $ is imposed and $\omega \neq 0$).
1163    
1164     \subsubsection{Splitting the geo-potential}
1165    
1166     For the purposes of initialization and reducing round-off errors, the model
1167     deals with perturbations from reference (or ``standard'') profiles. For
1168     example, the hydrostatic geopotential associated with the resting atmosphere
1169     is not dynamically relevant and can therefore be subtracted from the
1170     equations. The equations written in terms of perturbations are obtained by
1171     substituting the following definitions into the previous model equations:
1172     \begin{eqnarray}
1173     \theta &=&\theta _{o}+\theta ^{\prime } \label{eq:atmos-ref-prof-theta} \\
1174     \alpha &=&\alpha _{o}+\alpha ^{\prime } \label{eq:atmos-ref-prof-alpha} \\
1175     \phi &=&\phi _{o}+\phi ^{\prime } \label{eq:atmos-ref-prof-phi}
1176     \end{eqnarray}
1177     The reference state (indicated by subscript ``0'') corresponds to
1178     horizontally homogeneous atmosphere at rest ($\theta _{o},\alpha _{o},\phi
1179     _{o}$) with surface pressure $p_{o}(x,y)$ that satisfies $\phi
1180     _{o}(p_{o})=g~Z_{topo}$, defined:
1181     \begin{eqnarray*}
1182     \theta _{o}(p) &=&f^{n}(p) \\
1183     \alpha _{o}(p) &=&\Pi _{p}\theta _{o} \\
1184     \phi _{o}(p) &=&\phi _{topo}-\int_{p_{0}}^{p}\alpha _{o}dp
1185     \end{eqnarray*}
1186     %\begin{eqnarray*}
1187     %\phi'_\alpha & = & \int^p_{p_o} (\alpha_o -\alpha) dp \\
1188     %\phi'_s(x,y,t) & = & \int_{p_o}^{p_s} \alpha dp
1189     %\end{eqnarray*}
1190    
1191     The final form of the HPE's in p coordinates is then:
1192     \begin{eqnarray}
1193 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1194 cnh 1.8 _{h}+\mathbf{\nabla }_{p}\phi ^{\prime } &=&\vec{\mathbf{\mathcal{F}}} \label{eq:atmos-prime} \\
1195 cnh 1.1 \frac{\partial \phi ^{\prime }}{\partial p}+\alpha ^{\prime } &=&0 \\
1196 adcroft 1.4 \mathbf{\nabla }_{p}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \omega }{
1197 cnh 1.1 \partial p} &=&0 \\
1198     \frac{\partial \Pi }{\partial p}\theta ^{\prime } &=&\alpha ^{\prime } \\
1199 cnh 1.8 \frac{D\theta }{Dt} &=&\frac{\mathcal{Q}}{\Pi }
1200 cnh 1.1 \end{eqnarray}
1201    
1202 cnh 1.13 % $Header: /u/u0/gcmpack/mitgcmdoc/part1/manual.tex,v 1.12 2001/11/19 14:33:44 cnh Exp $
1203 cnh 1.1 % $Name: $
1204    
1205     \section{Appendix OCEAN}
1206    
1207     \subsection{Equations of motion for the ocean}
1208    
1209     We review here the method by which the standard (Boussinesq, incompressible)
1210     HPE's for the ocean written in z-coordinates are obtained. The
1211     non-Boussinesq equations for oceanic motion are:
1212     \begin{eqnarray}
1213 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1214 cnh 1.1 _{h}+\frac{1}{\rho }\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}} \\
1215     \epsilon _{nh}\frac{Dw}{Dt}+g+\frac{1}{\rho }\frac{\partial p}{\partial z}
1216     &=&\epsilon _{nh}\mathcal{F}_{w} \\
1217 adcroft 1.4 \frac{1}{\rho }\frac{D\rho }{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}
1218 cnh 1.8 _{h}+\frac{\partial w}{\partial z} &=&0 \label{eq-zns-cont}\\
1219     \rho &=&\rho (\theta ,S,p) \label{eq-zns-eos}\\
1220     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-zns-heat}\\
1221     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-zns-salt}
1222     \label{eq:non-boussinesq}
1223 adcroft 1.4 \end{eqnarray}
1224 cnh 1.1 These equations permit acoustics modes, inertia-gravity waves,
1225 cnh 1.10 non-hydrostatic motions, a geostrophic (Rossby) mode and a thermohaline
1226 cnh 1.1 mode. As written, they cannot be integrated forward consistently - if we
1227     step $\rho $ forward in (\ref{eq-zns-cont}), the answer will not be
1228 adcroft 1.4 consistent with that obtained by stepping (\ref{eq-zns-heat}) and (\ref
1229 cnh 1.1 {eq-zns-salt}) and then using (\ref{eq-zns-eos}) to yield $\rho $. It is
1230     therefore necessary to manipulate the system as follows. Differentiating the
1231     EOS (equation of state) gives:
1232    
1233     \begin{equation}
1234     \frac{D\rho }{Dt}=\left. \frac{\partial \rho }{\partial \theta }\right|
1235     _{S,p}\frac{D\theta }{Dt}+\left. \frac{\partial \rho }{\partial S}\right|
1236     _{\theta ,p}\frac{DS}{Dt}+\left. \frac{\partial \rho }{\partial p}\right|
1237     _{\theta ,S}\frac{Dp}{Dt} \label{EOSexpansion}
1238     \end{equation}
1239    
1240     Note that $\frac{\partial \rho }{\partial p}=\frac{1}{c_{s}^{2}}$ is the
1241 cnh 1.8 reciprocal of the sound speed ($c_{s}$) squared. Substituting into \ref{eq-zns-cont} gives:
1242 cnh 1.1 \begin{equation}
1243 adcroft 1.4 \frac{1}{\rho c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{
1244 cnh 1.1 v}}+\partial _{z}w\approx 0 \label{eq-zns-pressure}
1245     \end{equation}
1246     where we have used an approximation sign to indicate that we have assumed
1247     adiabatic motion, dropping the $\frac{D\theta }{Dt}$ and $\frac{DS}{Dt}$.
1248     Replacing \ref{eq-zns-cont} with \ref{eq-zns-pressure} yields a system that
1249     can be explicitly integrated forward:
1250     \begin{eqnarray}
1251 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1252 cnh 1.1 _{h}+\frac{1}{\rho }\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1253     \label{eq-cns-hmom} \\
1254     \epsilon _{nh}\frac{Dw}{Dt}+g+\frac{1}{\rho }\frac{\partial p}{\partial z}
1255     &=&\epsilon _{nh}\mathcal{F}_{w} \label{eq-cns-hydro} \\
1256 adcroft 1.4 \frac{1}{\rho c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{
1257 cnh 1.1 v}}_{h}+\frac{\partial w}{\partial z} &=&0 \label{eq-cns-cont} \\
1258     \rho &=&\rho (\theta ,S,p) \label{eq-cns-eos} \\
1259     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-cns-heat} \\
1260     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-cns-salt}
1261     \end{eqnarray}
1262    
1263     \subsubsection{Compressible z-coordinate equations}
1264    
1265     Here we linearize the acoustic modes by replacing $\rho $ with $\rho _{o}(z)$
1266     wherever it appears in a product (ie. non-linear term) - this is the
1267     `Boussinesq assumption'. The only term that then retains the full variation
1268     in $\rho $ is the gravitational acceleration:
1269     \begin{eqnarray}
1270 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1271 cnh 1.1 _{h}+\frac{1}{\rho _{o}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1272     \label{eq-zcb-hmom} \\
1273 adcroft 1.4 \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{o}}+\frac{1}{\rho _{o}}
1274 cnh 1.1 \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1275     \label{eq-zcb-hydro} \\
1276 adcroft 1.4 \frac{1}{\rho _{o}c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{
1277 cnh 1.1 \mathbf{v}}_{h}+\frac{\partial w}{\partial z} &=&0 \label{eq-zcb-cont} \\
1278     \rho &=&\rho (\theta ,S,p) \label{eq-zcb-eos} \\
1279     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-zcb-heat} \\
1280     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-zcb-salt}
1281     \end{eqnarray}
1282     These equations still retain acoustic modes. But, because the
1283 adcroft 1.4 ``compressible'' terms are linearized, the pressure equation \ref
1284 cnh 1.1 {eq-zcb-cont} can be integrated implicitly with ease (the time-dependent
1285     term appears as a Helmholtz term in the non-hydrostatic pressure equation).
1286     These are the \emph{truly} compressible Boussinesq equations. Note that the
1287     EOS must have the same pressure dependency as the linearized pressure term,
1288 adcroft 1.4 ie. $\left. \frac{\partial \rho }{\partial p}\right| _{\theta ,S}=\frac{1}{
1289 cnh 1.1 c_{s}^{2}}$, for consistency.
1290    
1291     \subsubsection{`Anelastic' z-coordinate equations}
1292    
1293     The anelastic approximation filters the acoustic mode by removing the
1294 adcroft 1.4 time-dependency in the continuity (now pressure-) equation (\ref{eq-zcb-cont}
1295     ). This could be done simply by noting that $\frac{Dp}{Dt}\approx -g\rho _{o}
1296 cnh 1.1 \frac{Dz}{Dt}=-g\rho _{o}w$, but this leads to an inconsistency between
1297     continuity and EOS. A better solution is to change the dependency on
1298     pressure in the EOS by splitting the pressure into a reference function of
1299     height and a perturbation:
1300     \begin{equation*}
1301 cnh 1.2 \rho =\rho (\theta ,S,p_{o}(z)+\epsilon _{s}p^{\prime })
1302 cnh 1.1 \end{equation*}
1303     Remembering that the term $\frac{Dp}{Dt}$ in continuity comes from
1304     differentiating the EOS, the continuity equation then becomes:
1305     \begin{equation*}
1306 adcroft 1.4 \frac{1}{\rho _{o}c_{s}^{2}}\left( \frac{Dp_{o}}{Dt}+\epsilon _{s}\frac{
1307     Dp^{\prime }}{Dt}\right) +\mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+
1308 cnh 1.2 \frac{\partial w}{\partial z}=0
1309 cnh 1.1 \end{equation*}
1310     If the time- and space-scales of the motions of interest are longer than
1311 adcroft 1.4 those of acoustic modes, then $\frac{Dp^{\prime }}{Dt}<<(\frac{Dp_{o}}{Dt},
1312 cnh 1.1 \mathbf{\nabla }\cdot \vec{\mathbf{v}}_{h})$ in the continuity equations and
1313 adcroft 1.4 $\left. \frac{\partial \rho }{\partial p}\right| _{\theta ,S}\frac{
1314 cnh 1.1 Dp^{\prime }}{Dt}<<\left. \frac{\partial \rho }{\partial p}\right| _{\theta
1315     ,S}\frac{Dp_{o}}{Dt}$ in the EOS (\ref{EOSexpansion}). Thus we set $\epsilon
1316     _{s}=0$, removing the dependency on $p^{\prime }$ in the continuity equation
1317     and EOS. Expanding $\frac{Dp_{o}(z)}{Dt}=-g\rho _{o}w$ then leads to the
1318     anelastic continuity equation:
1319     \begin{equation}
1320 adcroft 1.4 \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z}-
1321 cnh 1.1 \frac{g}{c_{s}^{2}}w=0 \label{eq-za-cont1}
1322     \end{equation}
1323     A slightly different route leads to the quasi-Boussinesq continuity equation
1324 adcroft 1.4 where we use the scaling $\frac{\partial \rho ^{\prime }}{\partial t}+
1325     \mathbf{\nabla }_{3}\cdot \rho ^{\prime }\vec{\mathbf{v}}<<\mathbf{\nabla }
1326 cnh 1.1 _{3}\cdot \rho _{o}\vec{\mathbf{v}}$ yielding:
1327     \begin{equation}
1328 adcroft 1.4 \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{1}{\rho _{o}}\frac{
1329 cnh 1.1 \partial \left( \rho _{o}w\right) }{\partial z}=0 \label{eq-za-cont2}
1330     \end{equation}
1331     Equations \ref{eq-za-cont1} and \ref{eq-za-cont2} are in fact the same
1332     equation if:
1333     \begin{equation}
1334     \frac{1}{\rho _{o}}\frac{\partial \rho _{o}}{\partial z}=\frac{-g}{c_{s}^{2}}
1335     \end{equation}
1336     Again, note that if $\rho _{o}$ is evaluated from prescribed $\theta _{o}$
1337 adcroft 1.4 and $S_{o}$ profiles, then the EOS dependency on $p_{o}$ and the term $\frac{
1338 cnh 1.1 g}{c_{s}^{2}}$ in continuity should be referred to those same profiles. The
1339     full set of `quasi-Boussinesq' or `anelastic' equations for the ocean are
1340     then:
1341     \begin{eqnarray}
1342 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1343 cnh 1.1 _{h}+\frac{1}{\rho _{o}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1344     \label{eq-zab-hmom} \\
1345 adcroft 1.4 \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{o}}+\frac{1}{\rho _{o}}
1346 cnh 1.1 \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1347     \label{eq-zab-hydro} \\
1348 adcroft 1.4 \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{1}{\rho _{o}}\frac{
1349 cnh 1.1 \partial \left( \rho _{o}w\right) }{\partial z} &=&0 \label{eq-zab-cont} \\
1350     \rho &=&\rho (\theta ,S,p_{o}(z)) \label{eq-zab-eos} \\
1351     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-zab-heat} \\
1352     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-zab-salt}
1353     \end{eqnarray}
1354    
1355     \subsubsection{Incompressible z-coordinate equations}
1356    
1357     Here, the objective is to drop the depth dependence of $\rho _{o}$ and so,
1358     technically, to also remove the dependence of $\rho $ on $p_{o}$. This would
1359     yield the ``truly'' incompressible Boussinesq equations:
1360     \begin{eqnarray}
1361 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1362 cnh 1.1 _{h}+\frac{1}{\rho _{c}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1363     \label{eq-ztb-hmom} \\
1364 adcroft 1.4 \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{c}}+\frac{1}{\rho _{c}}
1365 cnh 1.1 \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1366     \label{eq-ztb-hydro} \\
1367     \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z}
1368     &=&0 \label{eq-ztb-cont} \\
1369     \rho &=&\rho (\theta ,S) \label{eq-ztb-eos} \\
1370     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-ztb-heat} \\
1371     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-ztb-salt}
1372     \end{eqnarray}
1373     where $\rho _{c}$ is a constant reference density of water.
1374    
1375     \subsubsection{Compressible non-divergent equations}
1376    
1377     The above ``incompressible'' equations are incompressible in both the flow
1378     and the density. In many oceanic applications, however, it is important to
1379     retain compressibility effects in the density. To do this we must split the
1380     density thus:
1381     \begin{equation*}
1382     \rho =\rho _{o}+\rho ^{\prime }
1383 adcroft 1.4 \end{equation*}
1384 cnh 1.1 We then assert that variations with depth of $\rho _{o}$ are unimportant
1385     while the compressible effects in $\rho ^{\prime }$ are:
1386     \begin{equation*}
1387     \rho _{o}=\rho _{c}
1388 adcroft 1.4 \end{equation*}
1389 cnh 1.1 \begin{equation*}
1390     \rho ^{\prime }=\rho (\theta ,S,p_{o}(z))-\rho _{o}
1391 adcroft 1.4 \end{equation*}
1392 cnh 1.1 This then yields what we can call the semi-compressible Boussinesq
1393     equations:
1394     \begin{eqnarray}
1395 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1396     _{h}+\frac{1}{\rho _{c}}\mathbf{\nabla }_{z}p^{\prime } &=&\vec{\mathbf{
1397 cnh 1.1 \mathcal{F}}} \label{eq:ocean-mom} \\
1398     \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho ^{\prime }}{\rho _{c}}+\frac{1}{\rho
1399     _{c}}\frac{\partial p^{\prime }}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1400     \label{eq:ocean-wmom} \\
1401     \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z}
1402     &=&0 \label{eq:ocean-cont} \\
1403     \rho ^{\prime } &=&\rho (\theta ,S,p_{o}(z))-\rho _{c} \label{eq:ocean-eos}
1404     \\
1405     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq:ocean-theta} \\
1406     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq:ocean-salt}
1407 adcroft 1.4 \end{eqnarray}
1408 cnh 1.1 Note that the hydrostatic pressure of the resting fluid, including that
1409     associated with $\rho _{c}$, is subtracted out since it has no effect on the
1410     dynamics.
1411    
1412     Though necessary, the assumptions that go into these equations are messy
1413     since we essentially assume a different EOS for the reference density and
1414     the perturbation density. Nevertheless, it is the hydrostatic ($\epsilon
1415     _{nh}=0$ form of these equations that are used throughout the ocean modeling
1416     community and referred to as the primitive equations (HPE).
1417    
1418 cnh 1.13 % $Header: /u/u0/gcmpack/mitgcmdoc/part1/manual.tex,v 1.12 2001/11/19 14:33:44 cnh Exp $
1419 cnh 1.1 % $Name: $
1420    
1421     \section{Appendix:OPERATORS}
1422    
1423     \subsection{Coordinate systems}
1424    
1425     \subsubsection{Spherical coordinates}
1426    
1427     In spherical coordinates, the velocity components in the zonal, meridional
1428     and vertical direction respectively, are given by (see Fig.2) :
1429    
1430     \begin{equation*}
1431 cnh 1.6 u=r\cos \varphi \frac{D\lambda }{Dt}
1432 cnh 1.1 \end{equation*}
1433    
1434     \begin{equation*}
1435 cnh 1.6 v=r\frac{D\varphi }{Dt}\qquad
1436 cnh 1.1 \end{equation*}
1437     $\qquad \qquad \qquad \qquad $
1438    
1439     \begin{equation*}
1440 cnh 1.2 \dot{r}=\frac{Dr}{Dt}
1441 cnh 1.1 \end{equation*}
1442    
1443 cnh 1.6 Here $\varphi $ is the latitude, $\lambda $ the longitude, $r$ the radial
1444 cnh 1.1 distance of the particle from the center of the earth, $\Omega $ is the
1445     angular speed of rotation of the Earth and $D/Dt$ is the total derivative.
1446    
1447     The `grad' ($\nabla $) and `div' ($\nabla $.) operators are defined by, in
1448     spherical coordinates:
1449    
1450     \begin{equation*}
1451 cnh 1.6 \nabla \equiv \left( \frac{1}{r\cos \varphi }\frac{\partial }{\partial \lambda }
1452     ,\frac{1}{r}\frac{\partial }{\partial \varphi },\frac{\partial }{\partial r}
1453 cnh 1.2 \right)
1454 cnh 1.1 \end{equation*}
1455    
1456     \begin{equation*}
1457 cnh 1.6 \nabla .v\equiv \frac{1}{r\cos \varphi }\left\{ \frac{\partial u}{\partial
1458     \lambda }+\frac{\partial }{\partial \varphi }\left( v\cos \varphi \right) \right\}
1459 cnh 1.2 +\frac{1}{r^{2}}\frac{\partial \left( r^{2}\dot{r}\right) }{\partial r}
1460 cnh 1.1 \end{equation*}
1461    
1462 adcroft 1.4 %tci%\end{document}

  ViewVC Help
Powered by ViewVC 1.1.22