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

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

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


Revision 1.4 - (hide annotations) (download) (as text)
Wed Apr 5 02:27:32 2006 UTC (19 years, 3 months ago) by edhill
Branch: MAIN
Changes since 1.3: +9 -9 lines
File MIME type: application/x-tex
refer to the model uniformly as "MITgcm" and treat it as a proper noun

1 edhill 1.4 % $Header: /u/gcmpack/manual/part1/manual_fromjm.tex,v 1.3 2001/10/15 19:34:28 adcroft Exp $
2 cnh 1.1 % $Name: $
3    
4     \documentclass[12pt]{book}
5     \usepackage{amsmath}
6     \usepackage{html}
7     \usepackage{epsfig}
8     \usepackage{graphics,subfigure}
9     \usepackage{array}
10     \usepackage{multirow}
11     \usepackage{fancyhdr}
12     \usepackage{psfrag}
13    
14     %TCIDATA{OutputFilter=Latex.dll}
15     %TCIDATA{LastRevised=Thursday, October 04, 2001 14:41:22}
16     %TCIDATA{<META NAME="GraphicsSave" CONTENT="32">}
17     %TCIDATA{Language=American English}
18    
19     \fancyhead{}
20     \fancyhead[LO]{\slshape \rightmark}
21     \fancyhead[RE]{\slshape \leftmark}
22     \fancyhead[RO,LE]{\thepage}
23     \fancyfoot[CO,CE]{\today}
24     \fancyfoot[RO,LE]{ }
25     \renewcommand{\headrulewidth}{0.4pt}
26     \renewcommand{\footrulewidth}{0.4pt}
27     \setcounter{secnumdepth}{3}
28     \input{tcilatex}
29    
30     \begin{document}
31    
32     \tableofcontents
33    
34    
35     % Section: Overview
36    
37 edhill 1.4 % $Header: /u/gcmpack/manual/part1/manual_fromjm.tex,v 1.3 2001/10/15 19:34:28 adcroft 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 adcroft 1.2 models - see fig
58 cnh 1.1 \marginpar{
59     Fig.1 One model}\ref{fig:onemodel}
60    
61 adcroft 1.2 %% CNHbegin
62     %notci%\input{part1/one_model_figure}
63     %% CNHend
64    
65 cnh 1.1 \item it has a non-hydrostatic capability and so can be used to study both
66 adcroft 1.2 small-scale and large scale processes - see fig
67 cnh 1.1 \marginpar{
68     Fig.2 All scales}\ref{fig:all-scales}
69    
70 adcroft 1.2 %% CNHbegin
71     %notci%\input{part1/all_scales_figure}
72     %% CNHend
73    
74 cnh 1.1 \item finite volume techniques are employed yielding an intuitive
75     discretization and support for the treatment of irregular geometries using
76 adcroft 1.2 orthogonal curvilinear grids and shaved cells - see fig
77 cnh 1.1 \marginpar{
78 adcroft 1.2 Fig.3 Finite volumes}\ref{fig:finite-volumes}
79    
80     %% CNHbegin
81     %notci%\input{part1/fvol_figure}
82     %% CNHend
83 cnh 1.1
84     \item tangent linear and adjoint counterparts are automatically maintained
85     along with the forward model, permitting sensitivity and optimization
86     studies.
87    
88     \item the model is developed to perform efficiently on a wide variety of
89     computational platforms.
90     \end{itemize}
91    
92     Key publications reporting on and charting the development of the model are
93     listed in an Appendix.
94    
95     We begin by briefly showing some of the results of the model in action to
96     give a feel for the wide range of problems that can be addressed using it.
97    
98 edhill 1.4 % $Header: /u/gcmpack/manual/part1/manual_fromjm.tex,v 1.3 2001/10/15 19:34:28 adcroft Exp $
99 cnh 1.1 % $Name: $
100    
101     \section{Illustrations of the model in action}
102    
103 edhill 1.4 MITgcm has been designed and used to model a wide range of phenomena,
104 cnh 1.1 from convection on the scale of meters in the ocean to the global pattern of
105     atmospheric winds - see fig.2\ref{fig:all-scales}. To give a flavor of the
106     kinds of problems the model has been used to study, we briefly describe some
107     of them here. A more detailed description of the underlying formulation,
108     numerical algorithm and implementation that lie behind these calculations is
109     given later. Indeed many of the illustrative examples shown below can be
110     easily reproduced: simply download the model (the minimum you need is a PC
111     running linux, together with a FORTRAN\ 77 compiler) and follow the examples
112     described in detail in the documentation.
113    
114     \subsection{Global atmosphere: `Held-Suarez' benchmark}
115    
116     A novel feature of MITgcm is its ability to simulate both atmospheric and
117     oceanographic flows at both small and large scales.
118    
119 adcroft 1.2 Fig.E1a.\ref{fig:eddy_cs} shows an instantaneous plot of the 500$mb$
120 cnh 1.1 temperature field obtained using the atmospheric isomorph of MITgcm run at
121     2.8$^{\circ }$ resolution on the cubed sphere. We see cold air over the pole
122     (blue) and warm air along an equatorial band (red). Fully developed
123     baroclinic eddies spawned in the northern hemisphere storm track are
124     evident. There are no mountains or land-sea contrast in this calculation,
125     but you can easily put them in. The model is driven by relaxation to a
126     radiative-convective equilibrium profile, following the description set out
127     in Held and Suarez; 1994 designed to test atmospheric hydrodynamical cores -
128     there are no mountains or land-sea contrast.
129    
130 adcroft 1.2 %% CNHbegin
131     %notci%\input{part1/cubic_eddies_figure}
132     %% CNHend
133    
134 cnh 1.1 As described in Adcroft (2001), a `cubed sphere' is used to discretize the
135     globe permitting a uniform gridding and obviated the need to fourier filter.
136     The `vector-invariant' form of MITgcm supports any orthogonal curvilinear
137     grid, of which the cubed sphere is just one of many choices.
138    
139     Fig.E1b shows the 5-year mean, zonally averaged potential temperature, zonal
140     wind and meridional overturning streamfunction from a 20-level version of
141     the model. It compares favorable with more conventional spatial
142     discretization approaches.
143    
144     A regular spherical lat-lon grid can also be used.
145    
146 adcroft 1.2 %% CNHbegin
147     %notci%\input{part1/hs_zave_u_figure}
148     %% CNHend
149    
150 cnh 1.1 \subsection{Ocean gyres}
151    
152     Baroclinic instability is a ubiquitous process in the ocean, as well as the
153     atmosphere. Ocean eddies play an important role in modifying the
154     hydrographic structure and current systems of the oceans. Coarse resolution
155     models of the oceans cannot resolve the eddy field and yield rather broad,
156     diffusive patterns of ocean currents. But if the resolution of our models is
157     increased until the baroclinic instability process is resolved, numerical
158     solutions of a different and much more realistic kind, can be obtained.
159    
160     Fig. ?.? shows the surface temperature and velocity field obtained from
161     MITgcm run at $\frac{1}{6}^{\circ }$ horizontal resolution on a $lat-lon$
162     grid in which the pole has been rotated by 90$^{\circ }$ on to the equator
163     (to avoid the converging of meridian in northern latitudes). 21 vertical
164     levels are used in the vertical with a `lopped cell' representation of
165     topography. The development and propagation of anomalously warm and cold
166     eddies can be clearly been seen in the Gulf Stream region. The transport of
167     warm water northward by the mean flow of the Gulf Stream is also clearly
168     visible.
169    
170 adcroft 1.2 %% CNHbegin
171     %notci%\input{part1/ocean_gyres_figure}
172     %% CNHend
173    
174    
175 cnh 1.1 \subsection{Global ocean circulation}
176    
177     Fig.E2a shows the pattern of ocean currents at the surface of a 4$^{\circ }$
178     global ocean model run with 15 vertical levels. Lopped cells are used to
179     represent topography on a regular $lat-lon$ grid extending from 70$^{\circ
180     }N $ to 70$^{\circ }S$. The model is driven using monthly-mean winds with
181     mixed boundary conditions on temperature and salinity at the surface. The
182     transfer properties of ocean eddies, convection and mixing is parameterized
183     in this model.
184    
185     Fig.E2b shows the meridional overturning circulation of the global ocean in
186     Sverdrups.
187    
188 adcroft 1.2 %%CNHbegin
189     %notci%\input{part1/global_circ_figure}
190     %%CNHend
191    
192 cnh 1.1 \subsection{Convection and mixing over topography}
193    
194     Dense plumes generated by localized cooling on the continental shelf of the
195     ocean may be influenced by rotation when the deformation radius is smaller
196     than the width of the cooling region. Rather than gravity plumes, the
197     mechanism for moving dense fluid down the shelf is then through geostrophic
198     eddies. The simulation shown in the figure (blue is cold dense fluid, red is
199     warmer, lighter fluid) employs the non-hydrostatic capability of MITgcm to
200     trigger convection by surface cooling. The cold, dense water falls down the
201     slope but is deflected along the slope by rotation. It is found that
202     entrainment in the vertical plane is reduced when rotational control is
203     strong, and replaced by lateral entrainment due to the baroclinic
204     instability of the along-slope current.
205    
206 adcroft 1.2 %%CNHbegin
207     %notci%\input{part1/convect_and_topo}
208     %%CNHend
209    
210 cnh 1.1 \subsection{Boundary forced internal waves}
211    
212     The unique ability of MITgcm to treat non-hydrostatic dynamics in the
213     presence of complex geometry makes it an ideal tool to study internal wave
214     dynamics and mixing in oceanic canyons and ridges driven by large amplitude
215     barotropic tidal currents imposed through open boundary conditions.
216    
217     Fig. ?.? shows the influence of cross-slope topographic variations on
218     internal wave breaking - the cross-slope velocity is in color, the density
219     contoured. The internal waves are excited by application of open boundary
220     conditions on the left.\ They propagate to the sloping boundary (represented
221     using MITgcm's finite volume spatial discretization) where they break under
222     nonhydrostatic dynamics.
223    
224 adcroft 1.2 %%CNHbegin
225     %notci%\input{part1/boundary_forced_waves}
226     %%CNHend
227    
228 cnh 1.1 \subsection{Parameter sensitivity using the adjoint of MITgcm}
229    
230     Forward and tangent linear counterparts of MITgcm are supported using an
231     `automatic adjoint compiler'. These can be used in parameter sensitivity and
232     data assimilation studies.
233    
234     As one example of application of the MITgcm adjoint, Fig.E4 maps the
235     gradient $\frac{\partial J}{\partial \mathcal{H}}$where $J$ is the magnitude
236 adcroft 1.2 of the overturning streamfunction shown in fig?.? at 40$^{\circ }$N and $
237 cnh 1.1 \mathcal{H}$ is the air-sea heat flux 100 years before. We see that $J$ is
238     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 adcroft 1.2 %%CNHbegin
243     %notci%\input{part1/adj_hf_ocean_figure}
244     %%CNHend
245    
246 cnh 1.1 \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     etc. Figure ?.? shows an estimate of the time-mean surface elevation of the
254     ocean obtained by bringing the model in to consistency with altimetric and
255     in-situ observations over the period 1992-1997.
256    
257 adcroft 1.2 %% CNHbegin
258     %notci%\input{part1/globes_figure}
259     %% CNHend
260    
261 cnh 1.1 \subsection{Ocean biogeochemical cycles}
262    
263     MITgcm is being used to study global biogeochemical cycles in the ocean. For
264     example one can study the effects of interannual changes in meteorological
265     forcing and upper ocean circulation on the fluxes of carbon dioxide and
266     oxygen between the ocean and atmosphere. The figure shows the annual air-sea
267     flux of oxygen and its relation to density outcrops in the southern oceans
268     from a single year of a global, interannually varying simulation.
269    
270 adcroft 1.2 %%CNHbegin
271     %notci%\input{part1/biogeo_figure}
272     %%CNHend
273 cnh 1.1
274     \subsection{Simulations of laboratory experiments}
275    
276     Figure ?.? shows MITgcm being used to simulate a laboratory experiment
277     enquiring in to the dynamics of the Antarctic Circumpolar Current (ACC). An
278     initially homogeneous tank of water ($1m$ in diameter) is driven from its
279     free surface by a rotating heated disk. The combined action of mechanical
280     and thermal forcing creates a lens of fluid which becomes baroclinically
281     unstable. The stratification and depth of penetration of the lens is
282     arrested by its instability in a process analogous to that whic sets the
283     stratification of the ACC.
284    
285 adcroft 1.2 %%CNHbegin
286     %notci%\input{part1/lab_figure}
287     %%CNHend
288    
289 edhill 1.4 % $Header: /u/gcmpack/manual/part1/manual_fromjm.tex,v 1.3 2001/10/15 19:34:28 adcroft Exp $
290 cnh 1.1 % $Name: $
291    
292     \section{Continuous equations in `r' coordinates}
293    
294     To render atmosphere and ocean models from one dynamical core we exploit
295     `isomorphisms' between equation sets that govern the evolution of the
296 adcroft 1.2 respective fluids - see fig.4
297 cnh 1.1 \marginpar{
298     Fig.4. Isomorphisms}. One system of hydrodynamical equations is written down
299     and encoded. The model variables have different interpretations depending on
300     whether the atmosphere or ocean is being studied. Thus, for example, the
301     vertical coordinate `$r$' is interpreted as pressure, $p$, if we are
302     modeling the atmosphere and height, $z$, if we are modeling the ocean.
303    
304 adcroft 1.2 %%CNHbegin
305     %notci%\input{part1/zandpcoord_figure.tex}
306     %%CNHend
307    
308 cnh 1.1 The state of the fluid at any time is characterized by the distribution of
309     velocity $\vec{\mathbf{v}}$, active tracers $\theta $ and $S$, a
310     `geopotential' $\phi $ and density $\rho =\rho (\theta ,S,p)$ which may
311     depend on $\theta $, $S$, and $p$. The equations that govern the evolution
312     of these fields, obtained by applying the laws of classical mechanics and
313     thermodynamics to a Boussinesq, Navier-Stokes fluid are, written in terms of
314 adcroft 1.2 a generic vertical coordinate, $r$, see fig.5
315 cnh 1.1 \marginpar{
316     Fig.5 The vertical coordinate of model}:
317    
318 adcroft 1.2 %%CNHbegin
319     %notci%\input{part1/vertcoord_figure.tex}
320     %%CNHend
321    
322 cnh 1.1 \begin{equation*}
323 adcroft 1.2 \frac{D\vec{\mathbf{v}_{h}}}{Dt}+\left( 2\vec{\Omega}\times \vec{\mathbf{v}}
324     \right) _{h}+\mathbf{\nabla }_{h}\phi =\mathcal{F}_{\vec{\mathbf{v}_{h}}}
325 cnh 1.1 \text{ horizontal mtm}
326     \end{equation*}
327    
328     \begin{equation*}
329 adcroft 1.2 \frac{D\dot{r}}{Dt}+\widehat{k}\cdot \left( 2\vec{\Omega}\times \vec{\mathbf{
330 cnh 1.1 v}}\right) +\frac{\partial \phi }{\partial r}+b=\mathcal{F}_{\dot{r}}\text{
331     vertical mtm}
332     \end{equation*}
333    
334     \begin{equation}
335 adcroft 1.2 \mathbf{\nabla }_{h}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \dot{r}}{
336 cnh 1.1 \partial r}=0\text{ continuity} \label{eq:continuous}
337     \end{equation}
338    
339     \begin{equation*}
340     b=b(\theta ,S,r)\text{ equation of state}
341     \end{equation*}
342    
343     \begin{equation*}
344     \frac{D\theta }{Dt}=\mathcal{Q}_{\theta }\text{ potential temperature}
345     \end{equation*}
346    
347     \begin{equation*}
348     \frac{DS}{Dt}=\mathcal{Q}_{S}\text{ humidity/salinity}
349     \end{equation*}
350    
351     Here:
352    
353     \begin{equation*}
354     r\text{ is the vertical coordinate}
355     \end{equation*}
356    
357     \begin{equation*}
358     \frac{D}{Dt}=\frac{\partial }{\partial t}+\vec{\mathbf{v}}\cdot \nabla \text{
359     is the total derivative}
360     \end{equation*}
361    
362     \begin{equation*}
363 adcroft 1.2 \mathbf{\nabla }=\mathbf{\nabla }_{h}+\widehat{k}\frac{\partial }{\partial r}
364 cnh 1.1 \text{ is the `grad' operator}
365     \end{equation*}
366 adcroft 1.2 with $\mathbf{\nabla }_{h}$ operating in the horizontal and $\widehat{k}
367 cnh 1.1 \frac{\partial }{\partial r}$ operating in the vertical, where $\widehat{k}$
368     is a unit vector in the vertical
369    
370     \begin{equation*}
371     t\text{ is time}
372     \end{equation*}
373    
374     \begin{equation*}
375     \vec{\mathbf{v}}=(u,v,\dot{r})=(\vec{\mathbf{v}}_{h},\dot{r})\text{ is the
376     velocity}
377     \end{equation*}
378    
379     \begin{equation*}
380     \phi \text{ is the `pressure'/`geopotential'}
381     \end{equation*}
382    
383     \begin{equation*}
384     \vec{\Omega}\text{ is the Earth's rotation}
385     \end{equation*}
386    
387     \begin{equation*}
388     b\text{ is the `buoyancy'}
389     \end{equation*}
390    
391     \begin{equation*}
392     \theta \text{ is potential temperature}
393     \end{equation*}
394    
395     \begin{equation*}
396     S\text{ is specific humidity in the atmosphere; salinity in the ocean}
397     \end{equation*}
398    
399     \begin{equation*}
400 adcroft 1.2 \mathcal{F}_{\vec{\mathbf{v}}}\text{ are forcing and dissipation of }\vec{
401 cnh 1.1 \mathbf{v}}
402     \end{equation*}
403    
404     \begin{equation*}
405     \mathcal{Q}_{\theta }\mathcal{\ }\text{are forcing and dissipation of }\theta
406     \end{equation*}
407    
408     \begin{equation*}
409     \mathcal{Q}_{S}\mathcal{\ }\text{are forcing and dissipation of }S
410     \end{equation*}
411    
412     The $\mathcal{F}^{\prime }s$ and $\mathcal{Q}^{\prime }s$ are provided by
413     extensive `physics' packages for atmosphere and ocean described in Chapter 6.
414    
415     \subsection{Kinematic Boundary conditions}
416    
417     \subsubsection{vertical}
418    
419     at fixed and moving $r$ surfaces we set (see fig.5):
420    
421     \begin{equation}
422     \dot{r}=0atr=R_{fixed}(x,y)\text{ (ocean bottom, top of the atmosphere)}
423     \label{eq:fixedbc}
424     \end{equation}
425    
426     \begin{equation}
427     \dot{r}=\frac{Dr}{Dt}atr=R_{moving}\text{ \
428     (oceansurface,bottomoftheatmosphere)} \label{eq:movingbc}
429     \end{equation}
430    
431     Here
432    
433     \begin{equation*}
434     R_{moving}=R_{o}+\eta
435     \end{equation*}
436     where $R_{o}(x,y)$ is the `$r-$value' (height or pressure, depending on
437     whether we are in the atmosphere or ocean) of the `moving surface' in the
438     resting fluid and $\eta $ is the departure from $R_{o}(x,y)$ in the presence
439     of motion.
440    
441     \subsubsection{horizontal}
442    
443     \begin{equation}
444     \vec{\mathbf{v}}\cdot \vec{\mathbf{n}}=0 \label{eq:noflow}
445 adcroft 1.2 \end{equation}
446 cnh 1.1 where $\vec{\mathbf{n}}$ is the normal to a solid boundary.
447    
448     \subsection{Atmosphere}
449    
450     In the atmosphere, see fig.5, we interpret:
451    
452     \begin{equation}
453     r=p\text{ is the pressure} \label{eq:atmos-r}
454     \end{equation}
455    
456     \begin{equation}
457     \dot{r}=\frac{Dp}{Dt}=\omega \text{ is the vertical velocity in }p\text{
458     coordinates} \label{eq:atmos-omega}
459     \end{equation}
460    
461     \begin{equation}
462     \phi =g\,z\text{ is the geopotential height} \label{eq:atmos-phi}
463     \end{equation}
464    
465     \begin{equation}
466     b=\frac{\partial \Pi }{\partial p}\theta \text{ is the buoyancy}
467     \label{eq:atmos-b}
468     \end{equation}
469    
470     \begin{equation}
471     \theta =T(\frac{p_{c}}{p})^{\kappa }\text{ is potential temperature}
472     \label{eq:atmos-theta}
473     \end{equation}
474    
475     \begin{equation}
476     S=q,\text{ is the specific humidity} \label{eq:atmos-s}
477     \end{equation}
478     where
479    
480     \begin{equation*}
481     T\text{ is absolute temperature}
482 adcroft 1.2 \end{equation*}
483 cnh 1.1 \begin{equation*}
484     p\text{ is the pressure}
485 adcroft 1.2 \end{equation*}
486 cnh 1.1 \begin{eqnarray*}
487     &&z\text{ is the height of the pressure surface} \\
488     &&g\text{ is the acceleration due to gravity}
489     \end{eqnarray*}
490    
491     In the above the ideal gas law, $p=\rho RT$, has been expressed in terms of
492     the Exner function $\Pi (p)$ given by (see Appendix Atmosphere)
493     \begin{equation}
494     \Pi (p)=c_{p}(\frac{p}{p_{c}})^{\kappa } \label{eq:exner}
495 adcroft 1.2 \end{equation}
496 cnh 1.1 where $p_{c}$ is a reference pressure and $\kappa =R/c_{p}$ with $R$ the gas
497     constant and $c_{p}$ the specific heat of air at constant pressure.
498    
499     At the top of the atmosphere (which is `fixed' in our $r$ coordinate):
500    
501     \begin{equation*}
502     R_{fixed}=p_{top}=0
503     \end{equation*}
504     In a resting atmosphere the elevation of the mountains at the bottom is
505     given by
506     \begin{equation*}
507     R_{moving}=R_{o}(x,y)=p_{o}(x,y)
508     \end{equation*}
509     i.e. the (hydrostatic) pressure at the top of the mountains in a resting
510     atmosphere.
511    
512     The boundary conditions at top and bottom are given by:
513    
514     \begin{eqnarray}
515     &&\omega =0~\text{at }r=R_{fixed} \text{ (top of the atmosphere)}
516     \label{eq:fixed-bc-atmos} \\
517     \omega &=&\frac{Dp_{s}}{Dt}\text{; at }r=R_{moving}\text{ (bottom of the
518     atmosphere)} \label{eq:moving-bc-atmos}
519     \end{eqnarray}
520    
521     Then the (hydrostatic form of) eq(\ref{eq:continuous}) yields a consistent
522     set of atmospheric equations which, for convenience, are written out in $p$
523     coordinates in Appendix Atmosphere - see eqs(\ref{eq:atmos-prime}).
524    
525     \subsection{Ocean}
526    
527     In the ocean we interpret:
528     \begin{eqnarray}
529     r &=&z\text{ is the height} \label{eq:ocean-z} \\
530     \dot{r} &=&\frac{Dz}{Dt}=w\text{ is the vertical velocity}
531     \label{eq:ocean-w} \\
532     \phi &=&\frac{p}{\rho _{c}}\text{ is the pressure} \label{eq:ocean-p} \\
533     b(\theta ,S,r) &=&\frac{g}{\rho _{c}}\left( \rho (\theta ,S,r)-\rho
534     _{c}\right) \text{ is the buoyancy} \label{eq:ocean-b}
535     \end{eqnarray}
536     where $\rho _{c}$ is a fixed reference density of water and $g$ is the
537     acceleration due to gravity.\noindent
538    
539     In the above
540    
541     At the bottom of the ocean: $R_{fixed}(x,y)=-H(x,y)$.
542    
543     The surface of the ocean is given by: $R_{moving}=\eta $
544    
545 adcroft 1.2 The position of the resting free surface of the ocean is given by $
546 cnh 1.1 R_{o}=Z_{o}=0$.
547    
548     Boundary conditions are:
549    
550     \begin{eqnarray}
551     w &=&0~\text{at }r=R_{fixed}\text{ (ocean bottom)} \label{eq:fixed-bc-ocean}
552     \\
553 adcroft 1.2 w &=&\frac{D\eta }{Dt}\text{ at }r=R_{moving}=\eta \text{ (ocean surface)
554 cnh 1.1 \label{eq:moving-bc-ocean}}
555     \end{eqnarray}
556     where $\eta $ is the elevation of the free surface.
557    
558     Then eq(\ref{eq:continuous}) yields a consistent set of oceanic equations
559     which, for convenience, are written out in $z$ coordinates in Appendix Ocean
560     - see eqs(\ref{eq:ocean-mom}) to (\ref{eq:ocean-salt}).
561    
562     \subsection{Hydrostatic, Quasi-hydrostatic, Quasi-nonhydrostatic and
563     Non-hydrostatic forms}
564    
565     Let us separate $\phi $ in to surface, hydrostatic and non-hydrostatic terms:
566    
567     \begin{equation}
568     \phi (x,y,r)=\phi _{s}(x,y)+\phi _{hyd}(x,y,r)+\phi _{nh}(x,y,r)
569     \label{eq:phi-split}
570 adcroft 1.2 \end{equation}
571 cnh 1.1 and write eq(\ref{incompressible}a,b) in the form:
572    
573     \begin{equation}
574     \frac{\partial \vec{\mathbf{v}_{h}}}{\partial t}+\mathbf{\nabla }_{h}\phi
575     _{s}+\mathbf{\nabla }_{h}\phi _{hyd}+\epsilon _{nh}\mathbf{\nabla }_{h}\phi
576     _{nh}=\vec{\mathbf{G}}_{\vec{v}_{h}} \label{eq:mom-h}
577     \end{equation}
578    
579     \begin{equation}
580     \frac{\partial \phi _{hyd}}{\partial r}=-b \label{eq:hydrostatic}
581     \end{equation}
582    
583     \begin{equation}
584 adcroft 1.2 \epsilon _{nh}\frac{\partial \dot{r}}{\partial t}+\frac{\partial \phi _{nh}}{
585 cnh 1.1 \partial r}=G_{\dot{r}} \label{eq:mom-w}
586     \end{equation}
587     Here $\epsilon _{nh}$ is a non-hydrostatic parameter.
588    
589 adcroft 1.2 The $\left( \vec{\mathbf{G}}_{\vec{v}},G_{\dot{r}}\right) $ in eq(\ref
590 cnh 1.1 {eq:mom-h}) and (\ref{eq:mom-w}) represent advective, metric and Coriolis
591 adcroft 1.2 terms in the momentum equations. In spherical coordinates they take the form
592     \footnote{
593 cnh 1.1 In the hydrostatic primitive equations (\textbf{HPE}) all underlined terms
594 adcroft 1.2 in (\ref{eq:gu-speherical}), (\ref{eq:gv-spherical}) and (\ref
595 cnh 1.1 {eq:gw-spherical}) are omitted; the singly-underlined terms are included in
596 adcroft 1.2 the quasi-hydrostatic model (\textbf{QH}). The fully non-hydrostatic model (
597 cnh 1.1 \textbf{NH}) includes all terms.} - see Marshall et al 1997a for a full
598     discussion:
599    
600     \begin{equation}
601     \left.
602     \begin{tabular}{l}
603     $G_{u}=-\vec{\mathbf{v}}.\nabla u$ \\
604     $-\left\{ \underline{\frac{u\dot{r}}{{r}}}-\frac{uv\tan lat}{{r}}\right\} $
605     \\
606     $-\left\{ -2\Omega v\sin lat+\underline{2\Omega \dot{r}\cos lat}\right\} $
607     \\
608 adcroft 1.2 $+\mathcal{F}_{u}$
609     \end{tabular}
610 cnh 1.1 \ \right\} \left\{
611     \begin{tabular}{l}
612     \textit{advection} \\
613     \textit{metric} \\
614     \textit{Coriolis} \\
615 adcroft 1.2 \textit{\ Forcing/Dissipation}
616     \end{tabular}
617 cnh 1.1 \ \right. \qquad \label{eq:gu-speherical}
618     \end{equation}
619    
620     \begin{equation}
621     \left.
622     \begin{tabular}{l}
623     $G_{v}=-\vec{\mathbf{v}}.\nabla v$ \\
624     $-\left\{ \underline{\frac{v\dot{r}}{{r}}}-\frac{u^{2}\tan lat}{{r}}\right\}
625     $ \\
626     $-\left\{ -2\Omega u\sin lat\right\} $ \\
627 adcroft 1.2 $+\mathcal{F}_{v}$
628     \end{tabular}
629 cnh 1.1 \ \right\} \left\{
630     \begin{tabular}{l}
631     \textit{advection} \\
632     \textit{metric} \\
633     \textit{Coriolis} \\
634 adcroft 1.2 \textit{\ Forcing/Dissipation}
635     \end{tabular}
636 cnh 1.1 \ \right. \qquad \label{eq:gv-spherical}
637 adcroft 1.2 \end{equation}
638 cnh 1.1 \qquad \qquad \qquad \qquad \qquad
639    
640     \begin{equation}
641     \left.
642     \begin{tabular}{l}
643     $G_{\dot{r}}=-\underline{\underline{\vec{\mathbf{v}}.\nabla \dot{r}}}$ \\
644     $+\left\{ \underline{\frac{u^{_{^{2}}}+v^{2}}{{r}}}\right\} $ \\
645     ${+}\underline{{2\Omega u\cos lat}}$ \\
646 adcroft 1.2 $\underline{\underline{\mathcal{F}_{\dot{r}}}}$
647     \end{tabular}
648 cnh 1.1 \ \right\} \left\{
649     \begin{tabular}{l}
650     \textit{advection} \\
651     \textit{metric} \\
652     \textit{Coriolis} \\
653 adcroft 1.2 \textit{\ Forcing/Dissipation}
654     \end{tabular}
655 cnh 1.1 \ \right. \label{eq:gw-spherical}
656 adcroft 1.2 \end{equation}
657 cnh 1.1 \qquad \qquad \qquad \qquad \qquad
658    
659 adcroft 1.2 In the above `${r}$' is the distance from the center of the earth and `$lat$
660 cnh 1.1 ' is latitude.
661    
662     Grad and div operators in spherical coordinates are defined in appendix
663 adcroft 1.2 OPERATORS.
664 cnh 1.1 \marginpar{
665     Fig.6 Spherical polar coordinate system.}
666    
667 adcroft 1.2 %%CNHbegin
668     %notci%\input{part1/sphere_coord_figure.tex}
669     %%CNHend
670    
671 cnh 1.1 \subsubsection{Shallow atmosphere approximation}
672    
673     Most models are based on the `hydrostatic primitive equations' (HPE's) in
674     which the vertical momentum equation is reduced to a statement of
675     hydrostatic balance and the `traditional approximation' is made in which the
676     Coriolis force is treated approximately and the shallow atmosphere
677 edhill 1.4 approximation is made. MITgcm need not make the `traditional
678 cnh 1.1 approximation'. To be able to support consistent non-hydrostatic forms the
679 adcroft 1.2 shallow atmosphere approximation can be relaxed - when dividing through by $
680 cnh 1.1 r $ in, for example, (\ref{eq:gu-speherical}), we do not replace $r$ by $a$,
681     the radius of the earth.
682    
683     \subsubsection{Hydrostatic and quasi-hydrostatic forms}
684    
685     These are discussed at length in Marshall et al (1997a).
686    
687     In the `hydrostatic primitive equations' (\textbf{HPE)} all the underlined
688     terms in Eqs. (\ref{eq:gu-speherical} $\rightarrow $\ \ref{eq:gw-spherical})
689     are neglected and `${r}$' is replaced by `$a$', the mean radius of the
690     earth. Once the pressure is found at one level - e.g. by inverting a 2-d
691     Elliptic equation for $\phi _{s}$ at $r=R_{moving}$ - the pressure can be
692 adcroft 1.2 computed at all other levels by integration of the hydrostatic relation, eq(
693 cnh 1.1 \ref{eq:hydrostatic}).
694    
695     In the `quasi-hydrostatic' equations (\textbf{QH)} strict balance between
696     gravity and vertical pressure gradients is not imposed. The $2\Omega u\cos
697     \phi $ Coriolis term are not neglected and are balanced by a non-hydrostatic
698 adcroft 1.2 contribution to the pressure field: only the terms underlined twice in Eqs. (
699 cnh 1.1 \ref{eq:gu-speherical}$\rightarrow $\ \ref{eq:gw-spherical}) are set to zero
700     and, simultaneously, the shallow atmosphere approximation is relaxed. In
701     \textbf{QH}\ \textit{all} the metric terms are retained and the full
702     variation of the radial position of a particle monitored. The \textbf{QH}\
703     vertical momentum equation (\ref{eq:mom-w}) becomes:
704    
705     \begin{equation*}
706     \frac{\partial \phi _{nh}}{\partial r}=2\Omega u\cos lat
707     \end{equation*}
708     making a small correction to the hydrostatic pressure.
709    
710     \textbf{QH} has good energetic credentials - they are the same as for
711     \textbf{HPE}. Importantly, however, it has the same angular momentum
712     principle as the full non-hydrostatic model (\textbf{NH)} - see Marshall
713     et.al., 1997a. As in \textbf{HPE }only a 2-d elliptic problem need be solved.
714    
715     \subsubsection{Non-hydrostatic and quasi-nonhydrostatic forms}
716    
717     The MIT model presently supports a full non-hydrostatic ocean isomorph, but
718     only a quasi-non-hydrostatic atmospheric isomorph.
719    
720     \paragraph{Non-hydrostatic Ocean}
721    
722 adcroft 1.2 In the non-hydrostatic ocean model all terms in equations Eqs.(\ref
723 cnh 1.1 {eq:gu-speherical} $\rightarrow $\ \ref{eq:gw-spherical}) are retained. A
724     three dimensional elliptic equation must be solved subject to Neumann
725     boundary conditions (see below). It is important to note that use of the
726     full \textbf{NH} does not admit any new `fast' waves in to the system - the
727     incompressible condition eq(\ref{eq:continuous})c has already filtered out
728     acoustic modes. It does, however, ensure that the gravity waves are treated
729     accurately with an exact dispersion relation. The \textbf{NH} set has a
730     complete angular momentum principle and consistent energetics - see White
731     and Bromley, 1995; Marshall et.al.\ 1997a.
732    
733     \paragraph{Quasi-nonhydrostatic Atmosphere}
734    
735 adcroft 1.2 In the non-hydrostatic version of our atmospheric model we approximate $\dot{
736 cnh 1.1 r}$ in the vertical momentum eqs(\ref{eq:mom-w}) and (\ref{eq:gv-spherical})
737     (but only here) by:
738    
739     \begin{equation}
740     \dot{r}=\frac{Dp}{Dt}=\frac{1}{g}\frac{D\phi }{Dt} \label{eq:quasi-nh-w}
741 adcroft 1.2 \end{equation}
742 cnh 1.1 where $p_{hy}$ is the hydrostatic pressure.
743    
744     \subsubsection{Summary of equation sets supported by model}
745    
746     \paragraph{Atmosphere}
747    
748     Hydrostatic, and quasi-hydrostatic and quasi non-hydrostatic forms of the
749     compressible non-Boussinesq equations in $p-$coordinates are supported.
750    
751     \subparagraph{Hydrostatic and quasi-hydrostatic}
752    
753     The hydrostatic set is written out in $p-$coordinates in appendix Atmosphere
754     - see eq(\ref{eq:atmos-prime}).
755    
756     \subparagraph{Quasi-nonhydrostatic}
757    
758     A quasi-nonhydrostatic form is also supported.
759    
760     \paragraph{Ocean}
761    
762     \subparagraph{Hydrostatic and quasi-hydrostatic}
763    
764     Hydrostatic, and quasi-hydrostatic forms of the incompressible Boussinesq
765     equations in $z-$coordinates are supported.
766    
767     \subparagraph{Non-hydrostatic}
768    
769 adcroft 1.2 Non-hydrostatic forms of the incompressible Boussinesq equations in $z-$
770     coordinates are supported - see eqs(\ref{eq:ocean-mom}) to (\ref
771 cnh 1.1 {eq:ocean-salt}).
772    
773     \subsection{Solution strategy}
774    
775 adcroft 1.2 The method of solution employed in the \textbf{HPE}, \textbf{QH} and \textbf{
776     NH} models is summarized in Fig.7.
777 cnh 1.1 \marginpar{
778     Fig.7 Solution strategy} Under all dynamics, a 2-d elliptic equation is
779     first solved to find the surface pressure and the hydrostatic pressure at
780     any level computed from the weight of fluid above. Under \textbf{HPE} and
781     \textbf{QH} dynamics, the horizontal momentum equations are then stepped
782     forward and $\dot{r}$ found from continuity. Under \textbf{NH} dynamics a
783     3-d elliptic equation must be solved for the non-hydrostatic pressure before
784     stepping forward the horizontal momentum equations; $\dot{r}$ is found by
785     stepping forward the vertical momentum equation.
786    
787 adcroft 1.2 %%CNHbegin
788     %notci%\input{part1/solution_strategy_figure.tex}
789     %%CNHend
790    
791 cnh 1.1 There is no penalty in implementing \textbf{QH} over \textbf{HPE} except, of
792 adcroft 1.2 course, some complication that goes with the inclusion of $\cos \phi \ $
793 cnh 1.1 Coriolis terms and the relaxation of the shallow atmosphere approximation.
794     But this leads to negligible increase in computation. In \textbf{NH}, in
795     contrast, one additional elliptic equation - a three-dimensional one - must
796     be inverted for $p_{nh}$. However the `overhead' of the \textbf{NH} model is
797     essentially negligible in the hydrostatic limit (see detailed discussion in
798     Marshall et al, 1997) resulting in a non-hydrostatic algorithm that, in the
799     hydrostatic limit, is as computationally economic as the \textbf{HPEs}.
800    
801     \subsection{Finding the pressure field}
802    
803     Unlike the prognostic variables $u$, $v$, $w$, $\theta $ and $S$, the
804     pressure field must be obtained diagnostically. We proceed, as before, by
805     dividing the total (pressure/geo) potential in to three parts, a surface
806     part, $\phi _{s}(x,y)$, a hydrostatic part $\phi _{hyd}(x,y,r)$ and a
807     non-hydrostatic part $\phi _{nh}(x,y,r)$, as in (\ref{eq:phi-split}), and
808     writing the momentum equation as in (\ref{eq:mom-h}).
809    
810     \subsubsection{Hydrostatic pressure}
811    
812     Hydrostatic pressure is obtained by integrating (\ref{eq:hydrostatic})
813     vertically from $r=R_{o}$ where $\phi _{hyd}(r=R_{o})=0$, to yield:
814    
815     \begin{equation*}
816 adcroft 1.2 \int_{r}^{R_{o}}\frac{\partial \phi _{hyd}}{\partial r}dr=\left[ \phi _{hyd}
817 cnh 1.1 \right] _{r}^{R_{o}}=\int_{r}^{R_{o}}-bdr
818     \end{equation*}
819     and so
820    
821     \begin{equation}
822     \phi _{hyd}(x,y,r)=\int_{r}^{R_{o}}bdr \label{eq:hydro-phi}
823     \end{equation}
824    
825     The model can be easily modified to accommodate a loading term (e.g
826     atmospheric pressure pushing down on the ocean's surface) by setting:
827    
828     \begin{equation}
829     \phi _{hyd}(r=R_{o})=loading \label{eq:loading}
830     \end{equation}
831    
832     \subsubsection{Surface pressure}
833    
834 adcroft 1.2 The surface pressure equation can be obtained by integrating continuity, (
835 cnh 1.1 \ref{eq:continuous})c, vertically from $r=R_{fixed}$ to $r=R_{moving}$
836    
837     \begin{equation*}
838 adcroft 1.2 \int_{R_{fixed}}^{R_{moving}}\left( \mathbf{\nabla }_{h}\cdot \vec{\mathbf{v}
839 cnh 1.1 }_{h}+\partial _{r}\dot{r}\right) dr=0
840     \end{equation*}
841    
842     Thus:
843    
844     \begin{equation*}
845     \frac{\partial \eta }{\partial t}+\vec{\mathbf{v}}.\nabla \eta
846 adcroft 1.2 +\int_{R_{fixed}}^{R_{moving}}\mathbf{\nabla }_{h}\cdot \vec{\mathbf{v}}
847 cnh 1.1 _{h}dr=0
848     \end{equation*}
849 adcroft 1.2 where $\eta =R_{moving}-R_{o}$ is the free-surface $r$-anomaly in units of $
850 cnh 1.1 r $. The above can be rearranged to yield, using Leibnitz's theorem:
851    
852     \begin{equation}
853     \frac{\partial \eta }{\partial t}+\mathbf{\nabla }_{h}\cdot
854     \int_{R_{fixed}}^{R_{moving}}\vec{\mathbf{v}}_{h}dr=\text{source}
855     \label{eq:free-surface}
856 adcroft 1.2 \end{equation}
857 cnh 1.1 where we have incorporated a source term.
858    
859     Whether $\phi $ is pressure (ocean model, $p/\rho _{c}$) or geopotential
860     (atmospheric model), in (\ref{mtm-split}), the horizontal gradient term can
861     be written
862     \begin{equation}
863     \mathbf{\nabla }_{h}\phi _{s}=\mathbf{\nabla }_{h}\left( b_{s}\eta \right)
864     \label{eq:phi-surf}
865 adcroft 1.2 \end{equation}
866 cnh 1.1 where $b_{s}$ is the buoyancy at the surface.
867    
868 adcroft 1.2 In the hydrostatic limit ($\epsilon _{nh}=0$), Eqs(\ref{eq:mom-h}), (\ref
869 cnh 1.1 {eq:free-surface}) and (\ref{eq:phi-surf}) can be solved by inverting a 2-d
870     elliptic equation for $\phi _{s}$ as described in Chapter 2. Both `free
871     surface' and `rigid lid' approaches are available.
872    
873     \subsubsection{Non-hydrostatic pressure}
874    
875 adcroft 1.2 Taking the horizontal divergence of (\ref{hor-mtm}) and adding $\frac{
876 cnh 1.1 \partial }{\partial r}$ of (\ref{vertmtm}), invoking the continuity equation
877     (\ref{incompressible}), we deduce that:
878    
879     \begin{equation}
880 adcroft 1.2 \nabla _{3}^{2}\phi _{nh}=\nabla .\vec{\mathbf{G}}_{\vec{v}}-\left( \mathbf{
881     \nabla }_{h}^{2}\phi _{s}+\mathbf{\nabla }^{2}\phi _{hyd}\right) =\nabla .
882 cnh 1.1 \vec{\mathbf{F}} \label{eq:3d-invert}
883     \end{equation}
884    
885     For a given rhs this 3-d elliptic equation must be inverted for $\phi _{nh}$
886     subject to appropriate choice of boundary conditions. This method is usually
887     called \textit{The Pressure Method} [Harlow and Welch, 1965; Williams, 1969;
888     Potter, 1976]. In the hydrostatic primitive equations case (\textbf{HPE}),
889     the 3-d problem does not need to be solved.
890    
891     \paragraph{Boundary Conditions}
892    
893     We apply the condition of no normal flow through all solid boundaries - the
894     coasts (in the ocean) and the bottom:
895    
896     \begin{equation}
897     \vec{\mathbf{v}}.\widehat{n}=0 \label{nonormalflow}
898     \end{equation}
899     where $\widehat{n}$ is a vector of unit length normal to the boundary. The
900     kinematic condition (\ref{nonormalflow}) is also applied to the vertical
901 adcroft 1.2 velocity at $r=R_{moving}$. No-slip $\left( v_{T}=0\right) \ $or slip $
902 cnh 1.1 \left( \partial v_{T}/\partial n=0\right) \ $conditions are employed on the
903     tangential component of velocity, $v_{T}$, at all solid boundaries,
904     depending on the form chosen for the dissipative terms in the momentum
905     equations - see below.
906    
907     Eq.(\ref{nonormalflow}) implies, making use of (\ref{mtm-split}), that:
908    
909     \begin{equation}
910     \widehat{n}.\nabla \phi _{nh}=\widehat{n}.\vec{\mathbf{F}}
911     \label{eq:inhom-neumann-nh}
912     \end{equation}
913     where
914    
915     \begin{equation*}
916     \vec{\mathbf{F}}=\vec{\mathbf{G}}_{\vec{v}}-\left( \mathbf{\nabla }_{h}\phi
917     _{s}+\mathbf{\nabla }\phi _{hyd}\right)
918 adcroft 1.2 \end{equation*}
919 cnh 1.1 presenting inhomogeneous Neumann boundary conditions to the Elliptic problem
920     (\ref{eq:3d-invert}). As shown, for example, by Williams (1969), one can
921     exploit classical 3D potential theory and, by introducing an appropriately
922     chosen $\delta $-function sheet of `source-charge', replace the
923     inhomogeneous boundary condition on pressure by a homogeneous one. The
924 adcroft 1.2 source term $rhs$ in (\ref{eq:3d-invert}) is the divergence of the vector $
925     \vec{\mathbf{F}}.$ By simultaneously setting $
926 cnh 1.1 \begin{array}{l}
927 adcroft 1.2 \widehat{n}.\vec{\mathbf{F}}
928     \end{array}
929 cnh 1.1 =0$\ and $\widehat{n}.\nabla \phi _{nh}=0\ $on the boundary the following
930     self-consistent but simpler homogenized Elliptic problem is obtained:
931    
932     \begin{equation*}
933     \nabla ^{2}\phi _{nh}=\nabla .\widetilde{\vec{\mathbf{F}}}\qquad
934 adcroft 1.2 \end{equation*}
935 cnh 1.1 where $\widetilde{\vec{\mathbf{F}}}$ is a modified $\vec{\mathbf{F}}$ such
936 adcroft 1.2 that $\widetilde{\vec{\mathbf{F}}}.\widehat{n}=0$. As is implied by (\ref
937 cnh 1.1 {eq:inhom-neumann-nh}) the modified boundary condition becomes:
938    
939     \begin{equation}
940     \widehat{n}.\nabla \phi _{nh}=0 \label{eq:hom-neumann-nh}
941     \end{equation}
942    
943     If the flow is `close' to hydrostatic balance then the 3-d inversion
944     converges rapidly because $\phi _{nh}\ $is then only a small correction to
945     the hydrostatic pressure field (see the discussion in Marshall et al, a,b).
946    
947     The solution $\phi _{nh}\ $to (\ref{eq:3d-invert}) and (\ref{homneuman})
948     does not vanish at $r=R_{moving}$, and so refines the pressure there.
949    
950     \subsection{Forcing/dissipation}
951    
952     \subsubsection{Forcing}
953    
954     The forcing terms $\mathcal{F}$ on the rhs of the equations are provided by
955     `physics packages' described in detail in chapter ??.
956    
957     \subsubsection{Dissipation}
958    
959     \paragraph{Momentum}
960    
961     Many forms of momentum dissipation are available in the model. Laplacian and
962     biharmonic frictions are commonly used:
963    
964     \begin{equation}
965 adcroft 1.2 D_{V}=A_{h}\nabla _{h}^{2}v+A_{v}\frac{\partial ^{2}v}{\partial z^{2}}
966 cnh 1.1 +A_{4}\nabla _{h}^{4}v \label{eq:dissipation}
967     \end{equation}
968     where $A_{h}$ and $A_{v}\ $are (constant) horizontal and vertical viscosity
969     coefficients and $A_{4}\ $is the horizontal coefficient for biharmonic
970     friction. These coefficients are the same for all velocity components.
971    
972     \paragraph{Tracers}
973    
974     The mixing terms for the temperature and salinity equations have a similar
975     form to that of momentum except that the diffusion tensor can be
976 adcroft 1.2 non-diagonal and have varying coefficients. $\qquad $
977 cnh 1.1 \begin{equation}
978     D_{T,S}=\nabla .[\underline{\underline{K}}\nabla (T,S)]+K_{4}\nabla
979     _{h}^{4}(T,S) \label{eq:diffusion}
980     \end{equation}
981 adcroft 1.2 where $\underline{\underline{K}}\ $is the diffusion tensor and the $K_{4}\ $
982 cnh 1.1 horizontal coefficient for biharmonic diffusion. In the simplest case where
983     the subgrid-scale fluxes of heat and salt are parameterized with constant
984     horizontal and vertical diffusion coefficients, $\underline{\underline{K}}$,
985     reduces to a diagonal matrix with constant coefficients:
986    
987     \begin{equation}
988     \qquad \qquad \qquad \qquad K=\left(
989     \begin{array}{ccc}
990     K_{h} & 0 & 0 \\
991     0 & K_{h} & 0 \\
992 adcroft 1.2 0 & 0 & K_{v}
993 cnh 1.1 \end{array}
994     \right) \qquad \qquad \qquad \label{eq:diagonal-diffusion-tensor}
995     \end{equation}
996     where $K_{h}\ $and $K_{v}\ $are the horizontal and vertical diffusion
997     coefficients. These coefficients are the same for all tracers (temperature,
998     salinity ... ).
999    
1000     \subsection{Vector invariant form}
1001    
1002 adcroft 1.2 For some purposes it is advantageous to write momentum advection in eq(\ref
1003 cnh 1.1 {hor-mtm}) and (\ref{vertmtm}) in the (so-called) `vector invariant' form:
1004    
1005     \begin{equation}
1006 adcroft 1.2 \frac{D\vec{\mathbf{v}}}{Dt}=\frac{\partial \vec{\mathbf{v}}}{\partial t}
1007     +\left( \nabla \times \vec{\mathbf{v}}\right) \times \vec{\mathbf{v}}+\nabla
1008 cnh 1.1 \left[ \frac{1}{2}(\vec{\mathbf{v}}\cdot \vec{\mathbf{v}})\right]
1009     \label{eq:vi-identity}
1010 adcroft 1.2 \end{equation}
1011 cnh 1.1 This permits alternative numerical treatments of the non-linear terms based
1012     on their representation as a vorticity flux. Because gradients of coordinate
1013     vectors no longer appear on the rhs of (\ref{eq:vi-identity}), explicit
1014 adcroft 1.2 representation of the metric terms in (\ref{eq:gu-speherical}), (\ref
1015 cnh 1.1 {eq:gv-spherical}) and (\ref{eq:gw-spherical}), can be avoided: information
1016     about the geometry is contained in the areas and lengths of the volumes used
1017     to discretize the model.
1018    
1019     \subsection{Adjoint}
1020    
1021     Tangent linear and adjoint counterparts of the forward model and described
1022     in Chapter 5.
1023    
1024 edhill 1.4 % $Header: /u/gcmpack/manual/part1/manual_fromjm.tex,v 1.3 2001/10/15 19:34:28 adcroft Exp $
1025 cnh 1.1 % $Name: $
1026    
1027     \section{Appendix ATMOSPHERE}
1028    
1029     \subsection{Hydrostatic Primitive Equations for the Atmosphere in pressure
1030     coordinates}
1031    
1032     \label{sect-hpe-p}
1033    
1034     The hydrostatic primitive equations (HPEs) in p-coordinates are:
1035     \begin{eqnarray}
1036 adcroft 1.2 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1037 cnh 1.1 _{h}+\mathbf{\nabla }_{p}\phi &=&\vec{\mathbf{\mathcal{F}}}
1038     \label{eq:atmos-mom} \\
1039     \frac{\partial \phi }{\partial p}+\alpha &=&0 \label{eq-p-hydro-start} \\
1040 adcroft 1.2 \mathbf{\nabla }_{p}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \omega }{
1041 cnh 1.1 \partial p} &=&0 \label{eq:atmos-cont} \\
1042     p\alpha &=&RT \label{eq:atmos-eos} \\
1043     c_{v}\frac{DT}{Dt}+p\frac{D\alpha }{Dt} &=&\mathcal{Q} \label{eq:atmos-heat}
1044 adcroft 1.2 \end{eqnarray}
1045 cnh 1.1 where $\vec{\mathbf{v}}_{h}=(u,v,0)$ is the `horizontal' (on pressure
1046     surfaces) component of velocity,$\frac{D}{Dt}=\vec{\mathbf{v}}_{h}\cdot
1047     \mathbf{\nabla }_{p}+\omega \frac{\partial }{\partial p}$ is the total
1048     derivative, $f=2\Omega \sin lat$ is the Coriolis parameter, $\phi =gz$ is
1049 adcroft 1.2 the geopotential, $\alpha =1/\rho $ is the specific volume, $\omega =\frac{Dp
1050     }{Dt}$ is the vertical velocity in the $p-$coordinate. Equation(\ref
1051     {eq:atmos-heat}) is the first law of thermodynamics where internal energy $
1052     e=c_{v}T$, $T$ is temperature, $Q$ is the rate of heating per unit mass and $
1053 cnh 1.1 p\frac{D\alpha }{Dt}$ is the work done by the fluid in compressing.
1054    
1055     It is convenient to cast the heat equation in terms of potential temperature
1056     $\theta $ so that it looks more like a generic conservation law.
1057     Differentiating (\ref{eq:atmos-eos}) we get:
1058     \begin{equation*}
1059     p\frac{D\alpha }{Dt}+\alpha \frac{Dp}{Dt}=R\frac{DT}{Dt}
1060 adcroft 1.2 \end{equation*}
1061     which, when added to the heat equation (\ref{eq:atmos-heat}) and using $
1062 cnh 1.1 c_{p}=c_{v}+R$, gives:
1063     \begin{equation}
1064     c_{p}\frac{DT}{Dt}-\alpha \frac{Dp}{Dt}=\mathcal{Q}
1065     \label{eq-p-heat-interim}
1066 adcroft 1.2 \end{equation}
1067 cnh 1.1 Potential temperature is defined:
1068     \begin{equation}
1069     \theta =T(\frac{p_{c}}{p})^{\kappa } \label{eq:potential-temp}
1070 adcroft 1.2 \end{equation}
1071 cnh 1.1 where $p_{c}$ is a reference pressure and $\kappa =R/c_{p}$. For convenience
1072     we will make use of the Exner function $\Pi (p)$ which defined by:
1073     \begin{equation}
1074     \Pi (p)=c_{p}(\frac{p}{p_{c}})^{\kappa } \label{Exner}
1075 adcroft 1.2 \end{equation}
1076 cnh 1.1 The following relations will be useful and are easily expressed in terms of
1077     the Exner function:
1078     \begin{equation*}
1079     c_{p}T=\Pi \theta \;\;;\;\;\frac{\partial \Pi }{\partial p}=\frac{\kappa \Pi
1080 adcroft 1.2 }{p}\;\;;\;\;\alpha =\frac{\kappa \Pi \theta }{p}=\frac{\partial \ \Pi }{
1081     \partial p}\theta \;\;;\;\;\frac{D\Pi }{Dt}=\frac{\partial \Pi }{\partial p}
1082 cnh 1.1 \frac{Dp}{Dt}
1083 adcroft 1.2 \end{equation*}
1084 cnh 1.1 where $b=\frac{\partial \ \Pi }{\partial p}\theta $ is the buoyancy.
1085    
1086     The heat equation is obtained by noting that
1087     \begin{equation*}
1088     c_{p}\frac{DT}{Dt}=\frac{D(\Pi \theta )}{Dt}=\Pi \frac{D\theta }{Dt}+\theta
1089     \frac{D\Pi }{Dt}=\Pi \frac{D\theta }{Dt}+\alpha \frac{Dp}{Dt}
1090     \end{equation*}
1091     and on substituting into (\ref{eq-p-heat-interim}) gives:
1092     \begin{equation}
1093     \Pi \frac{D\theta }{Dt}=\mathcal{Q}
1094     \label{eq:potential-temperature-equation}
1095     \end{equation}
1096     which is in conservative form.
1097    
1098 adcroft 1.2 For convenience in the model we prefer to step forward (\ref
1099 cnh 1.1 {eq:potential-temperature-equation}) rather than (\ref{eq:atmos-heat}).
1100    
1101     \subsubsection{Boundary conditions}
1102    
1103     The upper and lower boundary conditions are :
1104     \begin{eqnarray}
1105     \mbox{at the top:}\;\;p=0 &&\text{, }\omega =\frac{Dp}{Dt}=0 \\
1106     \mbox{at the surface:}\;\;p=p_{s} &&\text{, }\phi =\phi _{topo}=g~Z_{topo}
1107     \label{eq:boundary-condition-atmosphere}
1108     \end{eqnarray}
1109     In $p$-coordinates, the upper boundary acts like a solid boundary ($\omega
1110     =0 $); in $z$-coordinates and the lower boundary is analogous to a free
1111     surface ($\phi $ is imposed and $\omega \neq 0$).
1112    
1113     \subsubsection{Splitting the geo-potential}
1114    
1115     For the purposes of initialization and reducing round-off errors, the model
1116     deals with perturbations from reference (or ``standard'') profiles. For
1117     example, the hydrostatic geopotential associated with the resting atmosphere
1118     is not dynamically relevant and can therefore be subtracted from the
1119     equations. The equations written in terms of perturbations are obtained by
1120     substituting the following definitions into the previous model equations:
1121     \begin{eqnarray}
1122     \theta &=&\theta _{o}+\theta ^{\prime } \label{eq:atmos-ref-prof-theta} \\
1123     \alpha &=&\alpha _{o}+\alpha ^{\prime } \label{eq:atmos-ref-prof-alpha} \\
1124     \phi &=&\phi _{o}+\phi ^{\prime } \label{eq:atmos-ref-prof-phi}
1125     \end{eqnarray}
1126     The reference state (indicated by subscript ``0'') corresponds to
1127     horizontally homogeneous atmosphere at rest ($\theta _{o},\alpha _{o},\phi
1128     _{o}$) with surface pressure $p_{o}(x,y)$ that satisfies $\phi
1129     _{o}(p_{o})=g~Z_{topo}$, defined:
1130     \begin{eqnarray*}
1131     \theta _{o}(p) &=&f^{n}(p) \\
1132     \alpha _{o}(p) &=&\Pi _{p}\theta _{o} \\
1133     \phi _{o}(p) &=&\phi _{topo}-\int_{p_{0}}^{p}\alpha _{o}dp
1134     \end{eqnarray*}
1135     %\begin{eqnarray*}
1136     %\phi'_\alpha & = & \int^p_{p_o} (\alpha_o -\alpha) dp \\
1137     %\phi'_s(x,y,t) & = & \int_{p_o}^{p_s} \alpha dp
1138     %\end{eqnarray*}
1139    
1140     The final form of the HPE's in p coordinates is then:
1141     \begin{eqnarray}
1142 adcroft 1.2 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1143 cnh 1.1 _{h}+\mathbf{\nabla }_{p}\phi ^{\prime } &=&\vec{\mathbf{\mathcal{F}}} \\
1144     \frac{\partial \phi ^{\prime }}{\partial p}+\alpha ^{\prime } &=&0 \\
1145 adcroft 1.2 \mathbf{\nabla }_{p}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \omega }{
1146 cnh 1.1 \partial p} &=&0 \\
1147     \frac{\partial \Pi }{\partial p}\theta ^{\prime } &=&\alpha ^{\prime } \\
1148     \frac{D\theta }{Dt} &=&\frac{\mathcal{Q}}{\Pi } \label{eq:atmos-prime}
1149     \end{eqnarray}
1150    
1151 edhill 1.4 % $Header: /u/gcmpack/manual/part1/manual_fromjm.tex,v 1.3 2001/10/15 19:34:28 adcroft Exp $
1152 cnh 1.1 % $Name: $
1153    
1154     \section{Appendix OCEAN}
1155    
1156     \subsection{Equations of motion for the ocean}
1157    
1158     We review here the method by which the standard (Boussinesq, incompressible)
1159     HPE's for the ocean written in z-coordinates are obtained. The
1160     non-Boussinesq equations for oceanic motion are:
1161     \begin{eqnarray}
1162 adcroft 1.2 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1163 cnh 1.1 _{h}+\frac{1}{\rho }\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}} \\
1164     \epsilon _{nh}\frac{Dw}{Dt}+g+\frac{1}{\rho }\frac{\partial p}{\partial z}
1165     &=&\epsilon _{nh}\mathcal{F}_{w} \\
1166 adcroft 1.2 \frac{1}{\rho }\frac{D\rho }{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}
1167 cnh 1.1 _{h}+\frac{\partial w}{\partial z} &=&0 \\
1168     \rho &=&\rho (\theta ,S,p) \\
1169     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \\
1170     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq:non-boussinesq}
1171 adcroft 1.2 \end{eqnarray}
1172 cnh 1.1 These equations permit acoustics modes, inertia-gravity waves,
1173     non-hydrostatic motions, a geostrophic (Rossby) mode and a thermo-haline
1174     mode. As written, they cannot be integrated forward consistently - if we
1175     step $\rho $ forward in (\ref{eq-zns-cont}), the answer will not be
1176 adcroft 1.2 consistent with that obtained by stepping (\ref{eq-zns-heat}) and (\ref
1177 cnh 1.1 {eq-zns-salt}) and then using (\ref{eq-zns-eos}) to yield $\rho $. It is
1178     therefore necessary to manipulate the system as follows. Differentiating the
1179     EOS (equation of state) gives:
1180    
1181     \begin{equation}
1182     \frac{D\rho }{Dt}=\left. \frac{\partial \rho }{\partial \theta }\right|
1183     _{S,p}\frac{D\theta }{Dt}+\left. \frac{\partial \rho }{\partial S}\right|
1184     _{\theta ,p}\frac{DS}{Dt}+\left. \frac{\partial \rho }{\partial p}\right|
1185     _{\theta ,S}\frac{Dp}{Dt} \label{EOSexpansion}
1186     \end{equation}
1187    
1188     Note that $\frac{\partial \rho }{\partial p}=\frac{1}{c_{s}^{2}}$ is the
1189 adcroft 1.2 reciprocal of the sound speed ($c_{s}$) squared. Substituting into \ref
1190 cnh 1.1 {eq-zns-cont} gives:
1191     \begin{equation}
1192 adcroft 1.2 \frac{1}{\rho c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{
1193 cnh 1.1 v}}+\partial _{z}w\approx 0 \label{eq-zns-pressure}
1194     \end{equation}
1195     where we have used an approximation sign to indicate that we have assumed
1196     adiabatic motion, dropping the $\frac{D\theta }{Dt}$ and $\frac{DS}{Dt}$.
1197     Replacing \ref{eq-zns-cont} with \ref{eq-zns-pressure} yields a system that
1198     can be explicitly integrated forward:
1199     \begin{eqnarray}
1200 adcroft 1.2 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1201 cnh 1.1 _{h}+\frac{1}{\rho }\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1202     \label{eq-cns-hmom} \\
1203     \epsilon _{nh}\frac{Dw}{Dt}+g+\frac{1}{\rho }\frac{\partial p}{\partial z}
1204     &=&\epsilon _{nh}\mathcal{F}_{w} \label{eq-cns-hydro} \\
1205 adcroft 1.2 \frac{1}{\rho c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{
1206 cnh 1.1 v}}_{h}+\frac{\partial w}{\partial z} &=&0 \label{eq-cns-cont} \\
1207     \rho &=&\rho (\theta ,S,p) \label{eq-cns-eos} \\
1208     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-cns-heat} \\
1209     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-cns-salt}
1210     \end{eqnarray}
1211    
1212     \subsubsection{Compressible z-coordinate equations}
1213    
1214     Here we linearize the acoustic modes by replacing $\rho $ with $\rho _{o}(z)$
1215     wherever it appears in a product (ie. non-linear term) - this is the
1216     `Boussinesq assumption'. The only term that then retains the full variation
1217     in $\rho $ is the gravitational acceleration:
1218     \begin{eqnarray}
1219 adcroft 1.2 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1220 cnh 1.1 _{h}+\frac{1}{\rho _{o}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1221     \label{eq-zcb-hmom} \\
1222 adcroft 1.2 \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{o}}+\frac{1}{\rho _{o}}
1223 cnh 1.1 \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1224     \label{eq-zcb-hydro} \\
1225 adcroft 1.2 \frac{1}{\rho _{o}c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{
1226 cnh 1.1 \mathbf{v}}_{h}+\frac{\partial w}{\partial z} &=&0 \label{eq-zcb-cont} \\
1227     \rho &=&\rho (\theta ,S,p) \label{eq-zcb-eos} \\
1228     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-zcb-heat} \\
1229     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-zcb-salt}
1230     \end{eqnarray}
1231     These equations still retain acoustic modes. But, because the
1232 adcroft 1.2 ``compressible'' terms are linearized, the pressure equation \ref
1233 cnh 1.1 {eq-zcb-cont} can be integrated implicitly with ease (the time-dependent
1234     term appears as a Helmholtz term in the non-hydrostatic pressure equation).
1235     These are the \emph{truly} compressible Boussinesq equations. Note that the
1236     EOS must have the same pressure dependency as the linearized pressure term,
1237 adcroft 1.2 ie. $\left. \frac{\partial \rho }{\partial p}\right| _{\theta ,S}=\frac{1}{
1238 cnh 1.1 c_{s}^{2}}$, for consistency.
1239    
1240     \subsubsection{`Anelastic' z-coordinate equations}
1241    
1242     The anelastic approximation filters the acoustic mode by removing the
1243 adcroft 1.2 time-dependency in the continuity (now pressure-) equation (\ref{eq-zcb-cont}
1244     ). This could be done simply by noting that $\frac{Dp}{Dt}\approx -g\rho _{o}
1245 cnh 1.1 \frac{Dz}{Dt}=-g\rho _{o}w$, but this leads to an inconsistency between
1246     continuity and EOS. A better solution is to change the dependency on
1247     pressure in the EOS by splitting the pressure into a reference function of
1248     height and a perturbation:
1249     \begin{equation*}
1250     \rho =\rho (\theta ,S,p_{o}(z)+\epsilon _{s}p^{\prime })
1251     \end{equation*}
1252     Remembering that the term $\frac{Dp}{Dt}$ in continuity comes from
1253     differentiating the EOS, the continuity equation then becomes:
1254     \begin{equation*}
1255 adcroft 1.2 \frac{1}{\rho _{o}c_{s}^{2}}\left( \frac{Dp_{o}}{Dt}+\epsilon _{s}\frac{
1256     Dp^{\prime }}{Dt}\right) +\mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+
1257 cnh 1.1 \frac{\partial w}{\partial z}=0
1258     \end{equation*}
1259     If the time- and space-scales of the motions of interest are longer than
1260 adcroft 1.2 those of acoustic modes, then $\frac{Dp^{\prime }}{Dt}<<(\frac{Dp_{o}}{Dt},
1261 cnh 1.1 \mathbf{\nabla }\cdot \vec{\mathbf{v}}_{h})$ in the continuity equations and
1262 adcroft 1.2 $\left. \frac{\partial \rho }{\partial p}\right| _{\theta ,S}\frac{
1263 cnh 1.1 Dp^{\prime }}{Dt}<<\left. \frac{\partial \rho }{\partial p}\right| _{\theta
1264     ,S}\frac{Dp_{o}}{Dt}$ in the EOS (\ref{EOSexpansion}). Thus we set $\epsilon
1265     _{s}=0$, removing the dependency on $p^{\prime }$ in the continuity equation
1266     and EOS. Expanding $\frac{Dp_{o}(z)}{Dt}=-g\rho _{o}w$ then leads to the
1267     anelastic continuity equation:
1268     \begin{equation}
1269 adcroft 1.2 \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z}-
1270 cnh 1.1 \frac{g}{c_{s}^{2}}w=0 \label{eq-za-cont1}
1271     \end{equation}
1272     A slightly different route leads to the quasi-Boussinesq continuity equation
1273 adcroft 1.2 where we use the scaling $\frac{\partial \rho ^{\prime }}{\partial t}+
1274     \mathbf{\nabla }_{3}\cdot \rho ^{\prime }\vec{\mathbf{v}}<<\mathbf{\nabla }
1275 cnh 1.1 _{3}\cdot \rho _{o}\vec{\mathbf{v}}$ yielding:
1276     \begin{equation}
1277 adcroft 1.2 \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{1}{\rho _{o}}\frac{
1278 cnh 1.1 \partial \left( \rho _{o}w\right) }{\partial z}=0 \label{eq-za-cont2}
1279     \end{equation}
1280     Equations \ref{eq-za-cont1} and \ref{eq-za-cont2} are in fact the same
1281     equation if:
1282     \begin{equation}
1283     \frac{1}{\rho _{o}}\frac{\partial \rho _{o}}{\partial z}=\frac{-g}{c_{s}^{2}}
1284     \end{equation}
1285     Again, note that if $\rho _{o}$ is evaluated from prescribed $\theta _{o}$
1286 adcroft 1.2 and $S_{o}$ profiles, then the EOS dependency on $p_{o}$ and the term $\frac{
1287 cnh 1.1 g}{c_{s}^{2}}$ in continuity should be referred to those same profiles. The
1288     full set of `quasi-Boussinesq' or `anelastic' equations for the ocean are
1289     then:
1290     \begin{eqnarray}
1291 adcroft 1.2 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1292 cnh 1.1 _{h}+\frac{1}{\rho _{o}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1293     \label{eq-zab-hmom} \\
1294 adcroft 1.2 \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{o}}+\frac{1}{\rho _{o}}
1295 cnh 1.1 \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1296     \label{eq-zab-hydro} \\
1297 adcroft 1.2 \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{1}{\rho _{o}}\frac{
1298 cnh 1.1 \partial \left( \rho _{o}w\right) }{\partial z} &=&0 \label{eq-zab-cont} \\
1299     \rho &=&\rho (\theta ,S,p_{o}(z)) \label{eq-zab-eos} \\
1300     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-zab-heat} \\
1301     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-zab-salt}
1302     \end{eqnarray}
1303    
1304     \subsubsection{Incompressible z-coordinate equations}
1305    
1306     Here, the objective is to drop the depth dependence of $\rho _{o}$ and so,
1307     technically, to also remove the dependence of $\rho $ on $p_{o}$. This would
1308     yield the ``truly'' incompressible Boussinesq equations:
1309     \begin{eqnarray}
1310 adcroft 1.2 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1311 cnh 1.1 _{h}+\frac{1}{\rho _{c}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1312     \label{eq-ztb-hmom} \\
1313 adcroft 1.2 \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{c}}+\frac{1}{\rho _{c}}
1314 cnh 1.1 \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1315     \label{eq-ztb-hydro} \\
1316     \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z}
1317     &=&0 \label{eq-ztb-cont} \\
1318     \rho &=&\rho (\theta ,S) \label{eq-ztb-eos} \\
1319     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-ztb-heat} \\
1320     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-ztb-salt}
1321     \end{eqnarray}
1322     where $\rho _{c}$ is a constant reference density of water.
1323    
1324     \subsubsection{Compressible non-divergent equations}
1325    
1326     The above ``incompressible'' equations are incompressible in both the flow
1327     and the density. In many oceanic applications, however, it is important to
1328     retain compressibility effects in the density. To do this we must split the
1329     density thus:
1330     \begin{equation*}
1331     \rho =\rho _{o}+\rho ^{\prime }
1332 adcroft 1.2 \end{equation*}
1333 cnh 1.1 We then assert that variations with depth of $\rho _{o}$ are unimportant
1334     while the compressible effects in $\rho ^{\prime }$ are:
1335     \begin{equation*}
1336     \rho _{o}=\rho _{c}
1337 adcroft 1.2 \end{equation*}
1338 cnh 1.1 \begin{equation*}
1339     \rho ^{\prime }=\rho (\theta ,S,p_{o}(z))-\rho _{o}
1340 adcroft 1.2 \end{equation*}
1341 cnh 1.1 This then yields what we can call the semi-compressible Boussinesq
1342     equations:
1343     \begin{eqnarray}
1344 adcroft 1.2 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1345     _{h}+\frac{1}{\rho _{c}}\mathbf{\nabla }_{z}p^{\prime } &=&\vec{\mathbf{
1346 cnh 1.1 \mathcal{F}}} \label{eq:ocean-mom} \\
1347     \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho ^{\prime }}{\rho _{c}}+\frac{1}{\rho
1348     _{c}}\frac{\partial p^{\prime }}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1349     \label{eq:ocean-wmom} \\
1350     \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z}
1351     &=&0 \label{eq:ocean-cont} \\
1352     \rho ^{\prime } &=&\rho (\theta ,S,p_{o}(z))-\rho _{c} \label{eq:ocean-eos}
1353     \\
1354     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq:ocean-theta} \\
1355     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq:ocean-salt}
1356 adcroft 1.2 \end{eqnarray}
1357 cnh 1.1 Note that the hydrostatic pressure of the resting fluid, including that
1358     associated with $\rho _{c}$, is subtracted out since it has no effect on the
1359     dynamics.
1360    
1361     Though necessary, the assumptions that go into these equations are messy
1362     since we essentially assume a different EOS for the reference density and
1363     the perturbation density. Nevertheless, it is the hydrostatic ($\epsilon
1364     _{nh}=0$ form of these equations that are used throughout the ocean modeling
1365     community and referred to as the primitive equations (HPE).
1366    
1367 edhill 1.4 % $Header: /u/gcmpack/manual/part1/manual_fromjm.tex,v 1.3 2001/10/15 19:34:28 adcroft Exp $
1368 cnh 1.1 % $Name: $
1369    
1370     \section{Appendix:OPERATORS}
1371    
1372     \subsection{Coordinate systems}
1373    
1374     \subsubsection{Spherical coordinates}
1375    
1376     In spherical coordinates, the velocity components in the zonal, meridional
1377     and vertical direction respectively, are given by (see Fig.2) :
1378    
1379     \begin{equation*}
1380     u=r\cos \phi \frac{D\lambda }{Dt}
1381     \end{equation*}
1382    
1383     \begin{equation*}
1384     v=r\frac{D\phi }{Dt}\qquad
1385     \end{equation*}
1386     $\qquad \qquad \qquad \qquad $
1387    
1388     \begin{equation*}
1389     \dot{r}=\frac{Dr}{Dt}
1390     \end{equation*}
1391    
1392     Here $\phi $ is the latitude, $\lambda $ the longitude, $r$ the radial
1393     distance of the particle from the center of the earth, $\Omega $ is the
1394     angular speed of rotation of the Earth and $D/Dt$ is the total derivative.
1395    
1396     The `grad' ($\nabla $) and `div' ($\nabla $.) operators are defined by, in
1397     spherical coordinates:
1398    
1399     \begin{equation*}
1400 adcroft 1.2 \nabla \equiv \left( \frac{1}{r\cos \phi }\frac{\partial }{\partial \lambda }
1401     ,\frac{1}{r}\frac{\partial }{\partial \phi },\frac{\partial }{\partial r}
1402 cnh 1.1 \right)
1403     \end{equation*}
1404    
1405     \begin{equation*}
1406     \nabla .v\equiv \frac{1}{r\cos \phi }\left\{ \frac{\partial u}{\partial
1407     \lambda }+\frac{\partial }{\partial \phi }\left( v\cos \phi \right) \right\}
1408     +\frac{1}{r^{2}}\frac{\partial \left( r^{2}\dot{r}\right) }{\partial r}
1409     \end{equation*}
1410    
1411     \end{document}

  ViewVC Help
Powered by ViewVC 1.1.22