/[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 1.28 - (hide annotations) (download) (as text)
Fri Aug 27 13:13:30 2010 UTC (14 years, 10 months ago) by jmc
Branch: MAIN
Changes since 1.27: +24 -24 lines
File MIME type: application/x-tex
changed for new directory path:
 s/input{part1\//input{s_overview\/text\//g
 /\\includegraphics.*{part1\//s/{part1\//{s_overview\/figs\//g

1 jmc 1.28 % $Header: /u/gcmpack/manual/s_overview/text/manual.tex,v 1.27 2008/01/17 21:28:22 jmc 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 jmc 1.28 % $Header: /u/gcmpack/manual/s_overview/text/manual.tex,v 1.27 2008/01/17 21:28:22 jmc Exp $
38 cnh 1.1 % $Name: $
39    
40 cnh 1.16 This document provides the reader with the information necessary to
41 cnh 1.1 carry out numerical experiments using MITgcm. It gives a comprehensive
42     description of the continuous equations on which the model is based, the
43     numerical algorithms the model employs and a description of the associated
44     program code. Along with the hydrodynamical kernel, physical and
45     biogeochemical parameterizations of key atmospheric and oceanic processes
46     are available. A number of examples illustrating the use of the model in
47     both process and general circulation studies of the atmosphere and ocean are
48     also presented.
49    
50 cnh 1.16 \section{Introduction}
51 afe 1.18 \begin{rawhtml}
52 afe 1.19 <!-- CMIREDIR:innovations: -->
53 afe 1.18 \end{rawhtml}
54    
55 cnh 1.16
56 cnh 1.1 MITgcm has a number of novel aspects:
57    
58     \begin{itemize}
59     \item it can be used to study both atmospheric and oceanic phenomena; one
60     hydrodynamical kernel is used to drive forward both atmospheric and oceanic
61 cnh 1.7 models - see fig \ref{fig:onemodel}
62 cnh 1.1
63 cnh 1.3 %% CNHbegin
64 jmc 1.28 \input{s_overview/text/one_model_figure}
65 cnh 1.3 %% CNHend
66    
67 cnh 1.1 \item it has a non-hydrostatic capability and so can be used to study both
68 cnh 1.7 small-scale and large scale processes - see fig \ref{fig:all-scales}
69 cnh 1.1
70 cnh 1.3 %% CNHbegin
71 jmc 1.28 \input{s_overview/text/all_scales_figure}
72 cnh 1.3 %% 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 cnh 1.7 orthogonal curvilinear grids and shaved cells - see fig \ref{fig:finite-volumes}
77 cnh 1.3
78     %% CNHbegin
79 jmc 1.28 \input{s_overview/text/fvol_figure}
80 cnh 1.3 %% CNHend
81 cnh 1.1
82     \item tangent linear and adjoint counterparts are automatically maintained
83     along with the forward model, permitting sensitivity and optimization
84     studies.
85    
86     \item the model is developed to perform efficiently on a wide variety of
87     computational platforms.
88     \end{itemize}
89    
90 jmc 1.27
91 cnh 1.16 Key publications reporting on and charting the development of the model are
92 jmc 1.27 \cite{hill:95,marshall:97a,marshall:97b,adcroft:97,mars-eta:98,adcroft:99,hill:99,maro-eta:99,adcroft:04a,adcroft:04b,marshall:04}
93     (an overview on the model formulation can also be found in \cite{adcroft:04c}):
94 cnh 1.12
95     \begin{verbatim}
96     Hill, C. and J. Marshall, (1995)
97     Application of a Parallel Navier-Stokes Model to Ocean Circulation in
98     Parallel Computational Fluid Dynamics
99     In Proceedings of Parallel Computational Fluid Dynamics: Implementations
100     and Results Using Parallel Computers, 545-552.
101     Elsevier Science B.V.: New York
102    
103     Marshall, J., C. Hill, L. Perelman, and A. Adcroft, (1997)
104 cnh 1.16 Hydrostatic, quasi-hydrostatic, and nonhydrostatic ocean modeling
105 cnh 1.12 J. Geophysical Res., 102(C3), 5733-5752.
106    
107     Marshall, J., A. Adcroft, C. Hill, L. Perelman, and C. Heisey, (1997)
108     A finite-volume, incompressible Navier Stokes model for studies of the ocean
109     on parallel computers,
110     J. Geophysical Res., 102(C3), 5753-5766.
111    
112     Adcroft, A.J., Hill, C.N. and J. Marshall, (1997)
113     Representation of topography by shaved cells in a height coordinate ocean
114     model
115     Mon Wea Rev, vol 125, 2293-2315
116    
117     Marshall, J., Jones, H. and C. Hill, (1998)
118     Efficient ocean modeling using non-hydrostatic algorithms
119     Journal of Marine Systems, 18, 115-134
120    
121     Adcroft, A., Hill C. and J. Marshall: (1999)
122     A new treatment of the Coriolis terms in C-grid models at both high and low
123     resolutions,
124     Mon. Wea. Rev. Vol 127, pages 1928-1936
125    
126     Hill, C, Adcroft,A., Jamous,D., and J. Marshall, (1999)
127     A Strategy for Terascale Climate Modeling.
128 cnh 1.14 In Proceedings of the Eighth ECMWF Workshop on the Use of Parallel Processors
129     in Meteorology, pages 406-425
130     World Scientific Publishing Co: UK
131 cnh 1.12
132     Marotzke, J, Giering,R., Zhang, K.Q., Stammer,D., Hill,C., and T.Lee, (1999)
133     Construction of the adjoint MIT ocean general circulation model and
134     application to Atlantic heat transport variability
135     J. Geophysical Res., 104(C12), 29,529-29,547.
136    
137     \end{verbatim}
138 cnh 1.1
139     We begin by briefly showing some of the results of the model in action to
140     give a feel for the wide range of problems that can be addressed using it.
141    
142 jmc 1.28 % $Header: /u/gcmpack/manual/s_overview/text/manual.tex,v 1.27 2008/01/17 21:28:22 jmc Exp $
143 cnh 1.1 % $Name: $
144    
145     \section{Illustrations of the model in action}
146    
147 edhill 1.24 MITgcm has been designed and used to model a wide range of phenomena,
148 cnh 1.1 from convection on the scale of meters in the ocean to the global pattern of
149 cnh 1.7 atmospheric winds - see figure \ref{fig:all-scales}. To give a flavor of the
150 cnh 1.1 kinds of problems the model has been used to study, we briefly describe some
151     of them here. A more detailed description of the underlying formulation,
152     numerical algorithm and implementation that lie behind these calculations is
153 cnh 1.2 given later. Indeed many of the illustrative examples shown below can be
154     easily reproduced: simply download the model (the minimum you need is a PC
155 cnh 1.10 running Linux, together with a FORTRAN\ 77 compiler) and follow the examples
156 cnh 1.2 described in detail in the documentation.
157 cnh 1.1
158     \subsection{Global atmosphere: `Held-Suarez' benchmark}
159 afe 1.18 \begin{rawhtml}
160 afe 1.19 <!-- CMIREDIR:atmospheric_example: -->
161 afe 1.18 \end{rawhtml}
162    
163    
164 cnh 1.1
165 cnh 1.7 A novel feature of MITgcm is its ability to simulate, using one basic algorithm,
166     both atmospheric and oceanographic flows at both small and large scales.
167 cnh 1.2
168 cnh 1.7 Figure \ref{fig:eddy_cs} shows an instantaneous plot of the 500$mb$
169 cnh 1.2 temperature field obtained using the atmospheric isomorph of MITgcm run at
170 edhill 1.25 $2.8^{\circ }$ resolution on the cubed sphere. We see cold air over the pole
171 cnh 1.2 (blue) and warm air along an equatorial band (red). Fully developed
172     baroclinic eddies spawned in the northern hemisphere storm track are
173     evident. There are no mountains or land-sea contrast in this calculation,
174     but you can easily put them in. The model is driven by relaxation to a
175     radiative-convective equilibrium profile, following the description set out
176     in Held and Suarez; 1994 designed to test atmospheric hydrodynamical cores -
177     there are no mountains or land-sea contrast.
178    
179 cnh 1.3 %% CNHbegin
180 jmc 1.28 \input{s_overview/text/cubic_eddies_figure}
181 cnh 1.3 %% CNHend
182    
183 cnh 1.2 As described in Adcroft (2001), a `cubed sphere' is used to discretize the
184 cnh 1.10 globe permitting a uniform griding and obviated the need to Fourier filter.
185 cnh 1.2 The `vector-invariant' form of MITgcm supports any orthogonal curvilinear
186     grid, of which the cubed sphere is just one of many choices.
187 cnh 1.1
188 cnh 1.7 Figure \ref{fig:hs_zave_u} shows the 5-year mean, zonally averaged zonal
189     wind from a 20-level configuration of
190 cnh 1.2 the model. It compares favorable with more conventional spatial
191 cnh 1.7 discretization approaches. The two plots show the field calculated using the
192     cube-sphere grid and the flow calculated using a regular, spherical polar
193     latitude-longitude grid. Both grids are supported within the model.
194 cnh 1.1
195 cnh 1.3 %% CNHbegin
196 jmc 1.28 \input{s_overview/text/hs_zave_u_figure}
197 cnh 1.3 %% CNHend
198    
199 cnh 1.2 \subsection{Ocean gyres}
200 afe 1.18 \begin{rawhtml}
201 afe 1.19 <!-- CMIREDIR:oceanic_example: -->
202 afe 1.18 \end{rawhtml}
203     \begin{rawhtml}
204 afe 1.19 <!-- CMIREDIR:ocean_gyres: -->
205 afe 1.18 \end{rawhtml}
206 cnh 1.1
207 cnh 1.2 Baroclinic instability is a ubiquitous process in the ocean, as well as the
208     atmosphere. Ocean eddies play an important role in modifying the
209     hydrographic structure and current systems of the oceans. Coarse resolution
210     models of the oceans cannot resolve the eddy field and yield rather broad,
211     diffusive patterns of ocean currents. But if the resolution of our models is
212     increased until the baroclinic instability process is resolved, numerical
213     solutions of a different and much more realistic kind, can be obtained.
214    
215 edhill 1.25 Figure \ref{fig:ocean-gyres} shows the surface temperature and
216     velocity field obtained from MITgcm run at $\frac{1}{6}^{\circ }$
217     horizontal resolution on a \textit{lat-lon} grid in which the pole has
218     been rotated by $90^{\circ }$ on to the equator (to avoid the
219     converging of meridian in northern latitudes). 21 vertical levels are
220     used in the vertical with a `lopped cell' representation of
221     topography. The development and propagation of anomalously warm and
222     cold eddies can be clearly seen in the Gulf Stream region. The
223     transport of warm water northward by the mean flow of the Gulf Stream
224     is also clearly visible.
225 cnh 1.1
226 cnh 1.3 %% CNHbegin
227 jmc 1.28 \input{s_overview/text/atl6_figure}
228 cnh 1.3 %% CNHend
229    
230    
231 cnh 1.1 \subsection{Global ocean circulation}
232 afe 1.18 \begin{rawhtml}
233 afe 1.19 <!-- CMIREDIR:global_ocean_circulation: -->
234 afe 1.18 \end{rawhtml}
235 cnh 1.1
236 edhill 1.25 Figure \ref{fig:large-scale-circ} (top) shows the pattern of ocean
237     currents at the surface of a $4^{\circ }$ global ocean model run with
238     15 vertical levels. Lopped cells are used to represent topography on a
239     regular \textit{lat-lon} grid extending from $70^{\circ }N$ to
240     $70^{\circ }S$. The model is driven using monthly-mean winds with
241     mixed boundary conditions on temperature and salinity at the surface.
242     The transfer properties of ocean eddies, convection and mixing is
243     parameterized in this model.
244 cnh 1.2
245 cnh 1.7 Figure \ref{fig:large-scale-circ} (bottom) shows the meridional overturning
246     circulation of the global ocean in Sverdrups.
247 cnh 1.2
248 cnh 1.3 %%CNHbegin
249 jmc 1.28 \input{s_overview/text/global_circ_figure}
250 cnh 1.3 %%CNHend
251    
252 cnh 1.2 \subsection{Convection and mixing over topography}
253 afe 1.18 \begin{rawhtml}
254 afe 1.19 <!-- CMIREDIR:mixing_over_topography: -->
255 afe 1.18 \end{rawhtml}
256    
257 cnh 1.2
258     Dense plumes generated by localized cooling on the continental shelf of the
259     ocean may be influenced by rotation when the deformation radius is smaller
260     than the width of the cooling region. Rather than gravity plumes, the
261     mechanism for moving dense fluid down the shelf is then through geostrophic
262 adcroft 1.9 eddies. The simulation shown in the figure \ref{fig:convect-and-topo}
263 cnh 1.7 (blue is cold dense fluid, red is
264 cnh 1.2 warmer, lighter fluid) employs the non-hydrostatic capability of MITgcm to
265     trigger convection by surface cooling. The cold, dense water falls down the
266     slope but is deflected along the slope by rotation. It is found that
267     entrainment in the vertical plane is reduced when rotational control is
268     strong, and replaced by lateral entrainment due to the baroclinic
269     instability of the along-slope current.
270 cnh 1.1
271 cnh 1.3 %%CNHbegin
272 jmc 1.28 \input{s_overview/text/convect_and_topo}
273 cnh 1.3 %%CNHend
274    
275 cnh 1.1 \subsection{Boundary forced internal waves}
276 afe 1.18 \begin{rawhtml}
277 afe 1.19 <!-- CMIREDIR:boundary_forced_internal_waves: -->
278 afe 1.18 \end{rawhtml}
279 cnh 1.1
280 cnh 1.2 The unique ability of MITgcm to treat non-hydrostatic dynamics in the
281     presence of complex geometry makes it an ideal tool to study internal wave
282     dynamics and mixing in oceanic canyons and ridges driven by large amplitude
283     barotropic tidal currents imposed through open boundary conditions.
284    
285 cnh 1.7 Fig. \ref{fig:boundary-forced-wave} shows the influence of cross-slope
286     topographic variations on
287 cnh 1.2 internal wave breaking - the cross-slope velocity is in color, the density
288     contoured. The internal waves are excited by application of open boundary
289 cnh 1.7 conditions on the left. They propagate to the sloping boundary (represented
290 cnh 1.2 using MITgcm's finite volume spatial discretization) where they break under
291     nonhydrostatic dynamics.
292    
293 cnh 1.3 %%CNHbegin
294 jmc 1.28 \input{s_overview/text/boundary_forced_waves}
295 cnh 1.3 %%CNHend
296    
297 cnh 1.2 \subsection{Parameter sensitivity using the adjoint of MITgcm}
298 afe 1.18 \begin{rawhtml}
299 afe 1.19 <!-- CMIREDIR:parameter_sensitivity: -->
300 afe 1.18 \end{rawhtml}
301 cnh 1.2
302     Forward and tangent linear counterparts of MITgcm are supported using an
303     `automatic adjoint compiler'. These can be used in parameter sensitivity and
304     data assimilation studies.
305    
306 edhill 1.25 As one example of application of the MITgcm adjoint, Figure
307     \ref{fig:hf-sensitivity} maps the gradient $\frac{\partial J}{\partial
308     \mathcal{H}}$where $J$ is the magnitude of the overturning
309     stream-function shown in figure \ref{fig:large-scale-circ} at
310     $60^{\circ }N$ and $ \mathcal{H}(\lambda,\varphi)$ is the mean, local
311     air-sea heat flux over a 100 year period. We see that $J$ is sensitive
312     to heat fluxes over the Labrador Sea, one of the important sources of
313     deep water for the thermohaline circulations. This calculation also
314 cnh 1.2 yields sensitivities to all other model parameters.
315    
316 cnh 1.3 %%CNHbegin
317 jmc 1.28 \input{s_overview/text/adj_hf_ocean_figure}
318 cnh 1.3 %%CNHend
319    
320 cnh 1.2 \subsection{Global state estimation of the ocean}
321 afe 1.18 \begin{rawhtml}
322 afe 1.19 <!-- CMIREDIR:global_state_estimation: -->
323 afe 1.18 \end{rawhtml}
324    
325 cnh 1.2
326     An important application of MITgcm is in state estimation of the global
327     ocean circulation. An appropriately defined `cost function', which measures
328     the departure of the model from observations (both remotely sensed and
329 cnh 1.10 in-situ) over an interval of time, is minimized by adjusting `control
330 cnh 1.2 parameters' such as air-sea fluxes, the wind field, the initial conditions
331 cnh 1.15 etc. Figure \ref{fig:assimilated-globes} shows the large scale planetary
332     circulation and a Hopf-Muller plot of Equatorial sea-surface height.
333     Both are obtained from assimilation bringing the model in to
334 cnh 1.7 consistency with altimetric and in-situ observations over the period
335 cnh 1.15 1992-1997.
336 cnh 1.2
337 cnh 1.3 %% CNHbegin
338 jmc 1.28 \input{s_overview/text/assim_figure}
339 cnh 1.3 %% CNHend
340    
341 cnh 1.2 \subsection{Ocean biogeochemical cycles}
342 afe 1.18 \begin{rawhtml}
343 afe 1.19 <!-- CMIREDIR:ocean_biogeo_cycles: -->
344 afe 1.18 \end{rawhtml}
345 cnh 1.2
346 edhill 1.25 MITgcm is being used to study global biogeochemical cycles in the
347     ocean. For example one can study the effects of interannual changes in
348     meteorological forcing and upper ocean circulation on the fluxes of
349     carbon dioxide and oxygen between the ocean and atmosphere. Figure
350     \ref{fig:biogeo} shows the annual air-sea flux of oxygen and its
351     relation to density outcrops in the southern oceans from a single year
352     of a global, interannually varying simulation. The simulation is run
353     at $1^{\circ}\times1^{\circ}$ resolution telescoping to
354     $\frac{1}{3}^{\circ}\times\frac{1}{3}^{\circ}$ in the tropics (not
355     shown).
356 cnh 1.2
357 cnh 1.3 %%CNHbegin
358 jmc 1.28 \input{s_overview/text/biogeo_figure}
359 cnh 1.3 %%CNHend
360 cnh 1.2
361     \subsection{Simulations of laboratory experiments}
362 afe 1.18 \begin{rawhtml}
363 afe 1.19 <!-- CMIREDIR:classroom_exp: -->
364 afe 1.18 \end{rawhtml}
365 cnh 1.2
366 cnh 1.7 Figure \ref{fig:lab-simulation} shows MITgcm being used to simulate a
367 edhill 1.17 laboratory experiment inquiring into the dynamics of the Antarctic Circumpolar Current (ACC). An
368 cnh 1.2 initially homogeneous tank of water ($1m$ in diameter) is driven from its
369     free surface by a rotating heated disk. The combined action of mechanical
370     and thermal forcing creates a lens of fluid which becomes baroclinically
371     unstable. The stratification and depth of penetration of the lens is
372 cnh 1.7 arrested by its instability in a process analogous to that which sets the
373 cnh 1.2 stratification of the ACC.
374 cnh 1.1
375 cnh 1.3 %%CNHbegin
376 jmc 1.28 \input{s_overview/text/lab_figure}
377 cnh 1.3 %%CNHend
378    
379 jmc 1.28 % $Header: /u/gcmpack/manual/s_overview/text/manual.tex,v 1.27 2008/01/17 21:28:22 jmc Exp $
380 cnh 1.1 % $Name: $
381    
382     \section{Continuous equations in `r' coordinates}
383 afe 1.18 \begin{rawhtml}
384 afe 1.19 <!-- CMIREDIR:z-p_isomorphism: -->
385 afe 1.18 \end{rawhtml}
386 cnh 1.1
387     To render atmosphere and ocean models from one dynamical core we exploit
388     `isomorphisms' between equation sets that govern the evolution of the
389 cnh 1.7 respective fluids - see figure \ref{fig:isomorphic-equations}.
390     One system of hydrodynamical equations is written down
391 cnh 1.1 and encoded. The model variables have different interpretations depending on
392     whether the atmosphere or ocean is being studied. Thus, for example, the
393     vertical coordinate `$r$' is interpreted as pressure, $p$, if we are
394 edhill 1.17 modeling the atmosphere (right hand side of figure \ref{fig:isomorphic-equations})
395     and height, $z$, if we are modeling the ocean (left hand side of figure
396 cnh 1.7 \ref{fig:isomorphic-equations}).
397 cnh 1.1
398 cnh 1.3 %%CNHbegin
399 jmc 1.28 \input{s_overview/text/zandpcoord_figure.tex}
400 cnh 1.3 %%CNHend
401    
402 cnh 1.1 The state of the fluid at any time is characterized by the distribution of
403     velocity $\vec{\mathbf{v}}$, active tracers $\theta $ and $S$, a
404     `geopotential' $\phi $ and density $\rho =\rho (\theta ,S,p)$ which may
405     depend on $\theta $, $S$, and $p$. The equations that govern the evolution
406     of these fields, obtained by applying the laws of classical mechanics and
407     thermodynamics to a Boussinesq, Navier-Stokes fluid are, written in terms of
408 cnh 1.7 a generic vertical coordinate, $r$, so that the appropriate
409     kinematic boundary conditions can be applied isomorphically
410     see figure \ref{fig:zandp-vert-coord}.
411 cnh 1.1
412 cnh 1.3 %%CNHbegin
413 jmc 1.28 \input{s_overview/text/vertcoord_figure.tex}
414 cnh 1.3 %%CNHend
415    
416 jmc 1.20 \begin{equation}
417 adcroft 1.4 \frac{D\vec{\mathbf{v}_{h}}}{Dt}+\left( 2\vec{\Omega}\times \vec{\mathbf{v}}
418     \right) _{h}+\mathbf{\nabla }_{h}\phi =\mathcal{F}_{\vec{\mathbf{v}_{h}}}
419 cnh 1.8 \text{ horizontal mtm} \label{eq:horizontal_mtm}
420 jmc 1.20 \end{equation}
421 cnh 1.1
422 cnh 1.8 \begin{equation}
423 adcroft 1.4 \frac{D\dot{r}}{Dt}+\widehat{k}\cdot \left( 2\vec{\Omega}\times \vec{\mathbf{
424 cnh 1.1 v}}\right) +\frac{\partial \phi }{\partial r}+b=\mathcal{F}_{\dot{r}}\text{
425 cnh 1.8 vertical mtm} \label{eq:vertical_mtm}
426     \end{equation}
427 cnh 1.1
428     \begin{equation}
429 adcroft 1.4 \mathbf{\nabla }_{h}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \dot{r}}{
430 cnh 1.8 \partial r}=0\text{ continuity} \label{eq:continuity}
431 cnh 1.1 \end{equation}
432    
433 cnh 1.8 \begin{equation}
434     b=b(\theta ,S,r)\text{ equation of state} \label{eq:equation_of_state}
435     \end{equation}
436 cnh 1.1
437 cnh 1.8 \begin{equation}
438 cnh 1.2 \frac{D\theta }{Dt}=\mathcal{Q}_{\theta }\text{ potential temperature}
439 cnh 1.8 \label{eq:potential_temperature}
440     \end{equation}
441 cnh 1.1
442 cnh 1.8 \begin{equation}
443 cnh 1.2 \frac{DS}{Dt}=\mathcal{Q}_{S}\text{ humidity/salinity}
444 adcroft 1.9 \label{eq:humidity_salt}
445 cnh 1.8 \end{equation}
446 cnh 1.1
447     Here:
448    
449     \begin{equation*}
450 cnh 1.2 r\text{ is the vertical coordinate}
451 cnh 1.1 \end{equation*}
452    
453     \begin{equation*}
454     \frac{D}{Dt}=\frac{\partial }{\partial t}+\vec{\mathbf{v}}\cdot \nabla \text{
455 cnh 1.2 is the total derivative}
456 cnh 1.1 \end{equation*}
457    
458     \begin{equation*}
459 adcroft 1.4 \mathbf{\nabla }=\mathbf{\nabla }_{h}+\widehat{k}\frac{\partial }{\partial r}
460 cnh 1.2 \text{ is the `grad' operator}
461 cnh 1.1 \end{equation*}
462 adcroft 1.4 with $\mathbf{\nabla }_{h}$ operating in the horizontal and $\widehat{k}
463 cnh 1.1 \frac{\partial }{\partial r}$ operating in the vertical, where $\widehat{k}$
464     is a unit vector in the vertical
465    
466     \begin{equation*}
467 cnh 1.2 t\text{ is time}
468 cnh 1.1 \end{equation*}
469    
470     \begin{equation*}
471     \vec{\mathbf{v}}=(u,v,\dot{r})=(\vec{\mathbf{v}}_{h},\dot{r})\text{ is the
472 cnh 1.2 velocity}
473 cnh 1.1 \end{equation*}
474    
475     \begin{equation*}
476 cnh 1.2 \phi \text{ is the `pressure'/`geopotential'}
477 cnh 1.1 \end{equation*}
478    
479     \begin{equation*}
480 cnh 1.2 \vec{\Omega}\text{ is the Earth's rotation}
481 cnh 1.1 \end{equation*}
482    
483     \begin{equation*}
484 cnh 1.2 b\text{ is the `buoyancy'}
485 cnh 1.1 \end{equation*}
486    
487     \begin{equation*}
488 cnh 1.2 \theta \text{ is potential temperature}
489 cnh 1.1 \end{equation*}
490    
491     \begin{equation*}
492 cnh 1.2 S\text{ is specific humidity in the atmosphere; salinity in the ocean}
493 cnh 1.1 \end{equation*}
494    
495     \begin{equation*}
496 adcroft 1.4 \mathcal{F}_{\vec{\mathbf{v}}}\text{ are forcing and dissipation of }\vec{
497 cnh 1.1 \mathbf{v}}
498     \end{equation*}
499    
500     \begin{equation*}
501 cnh 1.2 \mathcal{Q}_{\theta }\mathcal{\ }\text{are forcing and dissipation of }\theta
502 cnh 1.1 \end{equation*}
503    
504     \begin{equation*}
505     \mathcal{Q}_{S}\mathcal{\ }\text{are forcing and dissipation of }S
506     \end{equation*}
507    
508     The $\mathcal{F}^{\prime }s$ and $\mathcal{Q}^{\prime }s$ are provided by
509 cnh 1.7 `physics' and forcing packages for atmosphere and ocean. These are described
510     in later chapters.
511 cnh 1.1
512     \subsection{Kinematic Boundary conditions}
513    
514     \subsubsection{vertical}
515    
516 cnh 1.7 at fixed and moving $r$ surfaces we set (see figure \ref{fig:zandp-vert-coord}):
517 cnh 1.1
518     \begin{equation}
519 edhill 1.17 \dot{r}=0 \text{\ at\ } r=R_{fixed}(x,y)\text{ (ocean bottom, top of the atmosphere)}
520 cnh 1.1 \label{eq:fixedbc}
521     \end{equation}
522    
523     \begin{equation}
524 edhill 1.17 \dot{r}=\frac{Dr}{Dt} \text{\ at\ } r=R_{moving}\text{ \
525 cnh 1.10 (ocean surface,bottom of the atmosphere)} \label{eq:movingbc}
526 cnh 1.1 \end{equation}
527    
528     Here
529    
530     \begin{equation*}
531 cnh 1.2 R_{moving}=R_{o}+\eta
532 cnh 1.1 \end{equation*}
533     where $R_{o}(x,y)$ is the `$r-$value' (height or pressure, depending on
534     whether we are in the atmosphere or ocean) of the `moving surface' in the
535     resting fluid and $\eta $ is the departure from $R_{o}(x,y)$ in the presence
536     of motion.
537    
538     \subsubsection{horizontal}
539    
540     \begin{equation}
541     \vec{\mathbf{v}}\cdot \vec{\mathbf{n}}=0 \label{eq:noflow}
542 adcroft 1.4 \end{equation}
543 cnh 1.1 where $\vec{\mathbf{n}}$ is the normal to a solid boundary.
544    
545     \subsection{Atmosphere}
546    
547 cnh 1.7 In the atmosphere, (see figure \ref{fig:zandp-vert-coord}), we interpret:
548 cnh 1.1
549     \begin{equation}
550     r=p\text{ is the pressure} \label{eq:atmos-r}
551     \end{equation}
552    
553     \begin{equation}
554     \dot{r}=\frac{Dp}{Dt}=\omega \text{ is the vertical velocity in }p\text{
555     coordinates} \label{eq:atmos-omega}
556     \end{equation}
557    
558     \begin{equation}
559     \phi =g\,z\text{ is the geopotential height} \label{eq:atmos-phi}
560     \end{equation}
561    
562     \begin{equation}
563     b=\frac{\partial \Pi }{\partial p}\theta \text{ is the buoyancy}
564     \label{eq:atmos-b}
565     \end{equation}
566    
567     \begin{equation}
568     \theta =T(\frac{p_{c}}{p})^{\kappa }\text{ is potential temperature}
569     \label{eq:atmos-theta}
570     \end{equation}
571    
572     \begin{equation}
573     S=q,\text{ is the specific humidity} \label{eq:atmos-s}
574     \end{equation}
575     where
576    
577     \begin{equation*}
578     T\text{ is absolute temperature}
579 adcroft 1.4 \end{equation*}
580 cnh 1.1 \begin{equation*}
581     p\text{ is the pressure}
582 adcroft 1.4 \end{equation*}
583 cnh 1.1 \begin{eqnarray*}
584     &&z\text{ is the height of the pressure surface} \\
585     &&g\text{ is the acceleration due to gravity}
586     \end{eqnarray*}
587    
588     In the above the ideal gas law, $p=\rho RT$, has been expressed in terms of
589     the Exner function $\Pi (p)$ given by (see Appendix Atmosphere)
590     \begin{equation}
591     \Pi (p)=c_{p}(\frac{p}{p_{c}})^{\kappa } \label{eq:exner}
592 adcroft 1.4 \end{equation}
593 cnh 1.1 where $p_{c}$ is a reference pressure and $\kappa =R/c_{p}$ with $R$ the gas
594     constant and $c_{p}$ the specific heat of air at constant pressure.
595    
596     At the top of the atmosphere (which is `fixed' in our $r$ coordinate):
597    
598     \begin{equation*}
599 cnh 1.2 R_{fixed}=p_{top}=0
600 cnh 1.1 \end{equation*}
601     In a resting atmosphere the elevation of the mountains at the bottom is
602     given by
603     \begin{equation*}
604 cnh 1.2 R_{moving}=R_{o}(x,y)=p_{o}(x,y)
605 cnh 1.1 \end{equation*}
606     i.e. the (hydrostatic) pressure at the top of the mountains in a resting
607     atmosphere.
608    
609     The boundary conditions at top and bottom are given by:
610    
611     \begin{eqnarray}
612     &&\omega =0~\text{at }r=R_{fixed} \text{ (top of the atmosphere)}
613     \label{eq:fixed-bc-atmos} \\
614     \omega &=&\frac{Dp_{s}}{Dt}\text{; at }r=R_{moving}\text{ (bottom of the
615     atmosphere)} \label{eq:moving-bc-atmos}
616     \end{eqnarray}
617    
618 edhill 1.21 Then the (hydrostatic form of) equations
619     (\ref{eq:horizontal_mtm}-\ref{eq:humidity_salt}) yields a consistent
620     set of atmospheric equations which, for convenience, are written out
621     in $p$ coordinates in Appendix Atmosphere - see
622     eqs(\ref{eq:atmos-prime}).
623 cnh 1.1
624     \subsection{Ocean}
625    
626     In the ocean we interpret:
627     \begin{eqnarray}
628     r &=&z\text{ is the height} \label{eq:ocean-z} \\
629     \dot{r} &=&\frac{Dz}{Dt}=w\text{ is the vertical velocity}
630     \label{eq:ocean-w} \\
631     \phi &=&\frac{p}{\rho _{c}}\text{ is the pressure} \label{eq:ocean-p} \\
632     b(\theta ,S,r) &=&\frac{g}{\rho _{c}}\left( \rho (\theta ,S,r)-\rho
633     _{c}\right) \text{ is the buoyancy} \label{eq:ocean-b}
634     \end{eqnarray}
635     where $\rho _{c}$ is a fixed reference density of water and $g$ is the
636     acceleration due to gravity.\noindent
637    
638     In the above
639    
640     At the bottom of the ocean: $R_{fixed}(x,y)=-H(x,y)$.
641    
642     The surface of the ocean is given by: $R_{moving}=\eta $
643    
644 adcroft 1.4 The position of the resting free surface of the ocean is given by $
645 cnh 1.1 R_{o}=Z_{o}=0$.
646    
647     Boundary conditions are:
648    
649     \begin{eqnarray}
650     w &=&0~\text{at }r=R_{fixed}\text{ (ocean bottom)} \label{eq:fixed-bc-ocean}
651     \\
652 adcroft 1.4 w &=&\frac{D\eta }{Dt}\text{ at }r=R_{moving}=\eta \text{ (ocean surface)
653 cnh 1.1 \label{eq:moving-bc-ocean}}
654     \end{eqnarray}
655     where $\eta $ is the elevation of the free surface.
656    
657 adcroft 1.9 Then equations (\ref{eq:horizontal_mtm}-\ref{eq:humidity_salt}) yield a consistent set
658 cnh 1.8 of oceanic equations
659 cnh 1.1 which, for convenience, are written out in $z$ coordinates in Appendix Ocean
660     - see eqs(\ref{eq:ocean-mom}) to (\ref{eq:ocean-salt}).
661    
662     \subsection{Hydrostatic, Quasi-hydrostatic, Quasi-nonhydrostatic and
663     Non-hydrostatic forms}
664 afe 1.18 \begin{rawhtml}
665 afe 1.19 <!-- CMIREDIR:non_hydrostatic: -->
666 afe 1.18 \end{rawhtml}
667    
668 cnh 1.1
669     Let us separate $\phi $ in to surface, hydrostatic and non-hydrostatic terms:
670    
671     \begin{equation}
672     \phi (x,y,r)=\phi _{s}(x,y)+\phi _{hyd}(x,y,r)+\phi _{nh}(x,y,r)
673     \label{eq:phi-split}
674 adcroft 1.4 \end{equation}
675 jmc 1.20 %and write eq(\ref{eq:incompressible}) in the form:
676     % ^- this eq is missing (jmc) ; replaced with:
677     and write eq( \ref{eq:horizontal_mtm}) in the form:
678 cnh 1.1
679     \begin{equation}
680     \frac{\partial \vec{\mathbf{v}_{h}}}{\partial t}+\mathbf{\nabla }_{h}\phi
681     _{s}+\mathbf{\nabla }_{h}\phi _{hyd}+\epsilon _{nh}\mathbf{\nabla }_{h}\phi
682     _{nh}=\vec{\mathbf{G}}_{\vec{v}_{h}} \label{eq:mom-h}
683     \end{equation}
684    
685     \begin{equation}
686     \frac{\partial \phi _{hyd}}{\partial r}=-b \label{eq:hydrostatic}
687     \end{equation}
688    
689     \begin{equation}
690 adcroft 1.4 \epsilon _{nh}\frac{\partial \dot{r}}{\partial t}+\frac{\partial \phi _{nh}}{
691 cnh 1.1 \partial r}=G_{\dot{r}} \label{eq:mom-w}
692     \end{equation}
693     Here $\epsilon _{nh}$ is a non-hydrostatic parameter.
694    
695 adcroft 1.4 The $\left( \vec{\mathbf{G}}_{\vec{v}},G_{\dot{r}}\right) $ in eq(\ref
696 cnh 1.1 {eq:mom-h}) and (\ref{eq:mom-w}) represent advective, metric and Coriolis
697 adcroft 1.4 terms in the momentum equations. In spherical coordinates they take the form
698     \footnote{
699 cnh 1.1 In the hydrostatic primitive equations (\textbf{HPE}) all underlined terms
700 adcroft 1.4 in (\ref{eq:gu-speherical}), (\ref{eq:gv-spherical}) and (\ref
701 cnh 1.1 {eq:gw-spherical}) are omitted; the singly-underlined terms are included in
702 adcroft 1.4 the quasi-hydrostatic model (\textbf{QH}). The fully non-hydrostatic model (
703 cnh 1.1 \textbf{NH}) includes all terms.} - see Marshall et al 1997a for a full
704     discussion:
705    
706     \begin{equation}
707     \left.
708     \begin{tabular}{l}
709     $G_{u}=-\vec{\mathbf{v}}.\nabla u$ \\
710 cnh 1.6 $-\left\{ \underline{\frac{u\dot{r}}{{r}}}-\frac{uv\tan \varphi}{{r}}\right\} $
711 cnh 1.1 \\
712 cnh 1.6 $-\left\{ -2\Omega v\sin \varphi+\underline{2\Omega \dot{r}\cos \varphi}\right\} $
713 cnh 1.1 \\
714 adcroft 1.4 $+\mathcal{F}_{u}$
715     \end{tabular}
716 cnh 1.1 \ \right\} \left\{
717     \begin{tabular}{l}
718     \textit{advection} \\
719     \textit{metric} \\
720     \textit{Coriolis} \\
721 adcroft 1.4 \textit{\ Forcing/Dissipation}
722     \end{tabular}
723 cnh 1.2 \ \right. \qquad \label{eq:gu-speherical}
724 cnh 1.1 \end{equation}
725    
726     \begin{equation}
727     \left.
728     \begin{tabular}{l}
729     $G_{v}=-\vec{\mathbf{v}}.\nabla v$ \\
730 cnh 1.6 $-\left\{ \underline{\frac{v\dot{r}}{{r}}}-\frac{u^{2}\tan \varphi}{{r}}\right\}
731 cnh 1.1 $ \\
732 cnh 1.6 $-\left\{ -2\Omega u\sin \varphi \right\} $ \\
733 adcroft 1.4 $+\mathcal{F}_{v}$
734     \end{tabular}
735 cnh 1.1 \ \right\} \left\{
736     \begin{tabular}{l}
737     \textit{advection} \\
738     \textit{metric} \\
739     \textit{Coriolis} \\
740 adcroft 1.4 \textit{\ Forcing/Dissipation}
741     \end{tabular}
742 cnh 1.2 \ \right. \qquad \label{eq:gv-spherical}
743 adcroft 1.4 \end{equation}
744 cnh 1.2 \qquad \qquad \qquad \qquad \qquad
745 cnh 1.1
746     \begin{equation}
747     \left.
748     \begin{tabular}{l}
749     $G_{\dot{r}}=-\underline{\underline{\vec{\mathbf{v}}.\nabla \dot{r}}}$ \\
750     $+\left\{ \underline{\frac{u^{_{^{2}}}+v^{2}}{{r}}}\right\} $ \\
751 cnh 1.6 ${+}\underline{{2\Omega u\cos \varphi}}$ \\
752 adcroft 1.4 $\underline{\underline{\mathcal{F}_{\dot{r}}}}$
753     \end{tabular}
754 cnh 1.1 \ \right\} \left\{
755     \begin{tabular}{l}
756     \textit{advection} \\
757     \textit{metric} \\
758     \textit{Coriolis} \\
759 adcroft 1.4 \textit{\ Forcing/Dissipation}
760     \end{tabular}
761 cnh 1.2 \ \right. \label{eq:gw-spherical}
762 adcroft 1.4 \end{equation}
763 cnh 1.2 \qquad \qquad \qquad \qquad \qquad
764 cnh 1.1
765 cnh 1.6 In the above `${r}$' is the distance from the center of the earth and `$\varphi$
766 cnh 1.1 ' is latitude.
767    
768     Grad and div operators in spherical coordinates are defined in appendix
769 adcroft 1.4 OPERATORS.
770 cnh 1.1
771 cnh 1.3 %%CNHbegin
772 jmc 1.28 \input{s_overview/text/sphere_coord_figure.tex}
773 cnh 1.3 %%CNHend
774    
775 cnh 1.1 \subsubsection{Shallow atmosphere approximation}
776    
777 edhill 1.24 Most models are based on the `hydrostatic primitive equations' (HPE's)
778     in which the vertical momentum equation is reduced to a statement of
779     hydrostatic balance and the `traditional approximation' is made in
780     which the Coriolis force is treated approximately and the shallow
781     atmosphere approximation is made. MITgcm need not make the
782     `traditional approximation'. To be able to support consistent
783     non-hydrostatic forms the shallow atmosphere approximation can be
784     relaxed - when dividing through by $ r $ in, for example,
785     (\ref{eq:gu-speherical}), we do not replace $r$ by $a$, the radius of
786     the earth.
787 cnh 1.1
788     \subsubsection{Hydrostatic and quasi-hydrostatic forms}
789 cnh 1.7 \label{sec:hydrostatic_and_quasi-hydrostatic_forms}
790 cnh 1.1
791     These are discussed at length in Marshall et al (1997a).
792    
793     In the `hydrostatic primitive equations' (\textbf{HPE)} all the underlined
794     terms in Eqs. (\ref{eq:gu-speherical} $\rightarrow $\ \ref{eq:gw-spherical})
795     are neglected and `${r}$' is replaced by `$a$', the mean radius of the
796     earth. Once the pressure is found at one level - e.g. by inverting a 2-d
797     Elliptic equation for $\phi _{s}$ at $r=R_{moving}$ - the pressure can be
798 adcroft 1.4 computed at all other levels by integration of the hydrostatic relation, eq(
799 cnh 1.1 \ref{eq:hydrostatic}).
800    
801     In the `quasi-hydrostatic' equations (\textbf{QH)} strict balance between
802     gravity and vertical pressure gradients is not imposed. The $2\Omega u\cos
803 cnh 1.6 \varphi $ Coriolis term are not neglected and are balanced by a non-hydrostatic
804 adcroft 1.4 contribution to the pressure field: only the terms underlined twice in Eqs. (
805 cnh 1.1 \ref{eq:gu-speherical}$\rightarrow $\ \ref{eq:gw-spherical}) are set to zero
806     and, simultaneously, the shallow atmosphere approximation is relaxed. In
807     \textbf{QH}\ \textit{all} the metric terms are retained and the full
808     variation of the radial position of a particle monitored. The \textbf{QH}\
809     vertical momentum equation (\ref{eq:mom-w}) becomes:
810    
811     \begin{equation*}
812 cnh 1.6 \frac{\partial \phi _{nh}}{\partial r}=2\Omega u\cos \varphi
813 cnh 1.1 \end{equation*}
814     making a small correction to the hydrostatic pressure.
815    
816     \textbf{QH} has good energetic credentials - they are the same as for
817     \textbf{HPE}. Importantly, however, it has the same angular momentum
818     principle as the full non-hydrostatic model (\textbf{NH)} - see Marshall
819     et.al., 1997a. As in \textbf{HPE }only a 2-d elliptic problem need be solved.
820    
821     \subsubsection{Non-hydrostatic and quasi-nonhydrostatic forms}
822    
823 edhill 1.24 MITgcm presently supports a full non-hydrostatic ocean isomorph, but
824 cnh 1.1 only a quasi-non-hydrostatic atmospheric isomorph.
825    
826     \paragraph{Non-hydrostatic Ocean}
827    
828 adcroft 1.4 In the non-hydrostatic ocean model all terms in equations Eqs.(\ref
829 cnh 1.1 {eq:gu-speherical} $\rightarrow $\ \ref{eq:gw-spherical}) are retained. A
830     three dimensional elliptic equation must be solved subject to Neumann
831     boundary conditions (see below). It is important to note that use of the
832     full \textbf{NH} does not admit any new `fast' waves in to the system - the
833 cnh 1.8 incompressible condition eq(\ref{eq:continuity}) has already filtered out
834 cnh 1.1 acoustic modes. It does, however, ensure that the gravity waves are treated
835     accurately with an exact dispersion relation. The \textbf{NH} set has a
836     complete angular momentum principle and consistent energetics - see White
837     and Bromley, 1995; Marshall et.al.\ 1997a.
838    
839     \paragraph{Quasi-nonhydrostatic Atmosphere}
840    
841 adcroft 1.4 In the non-hydrostatic version of our atmospheric model we approximate $\dot{
842 cnh 1.1 r}$ in the vertical momentum eqs(\ref{eq:mom-w}) and (\ref{eq:gv-spherical})
843     (but only here) by:
844    
845     \begin{equation}
846     \dot{r}=\frac{Dp}{Dt}=\frac{1}{g}\frac{D\phi }{Dt} \label{eq:quasi-nh-w}
847 adcroft 1.4 \end{equation}
848 cnh 1.1 where $p_{hy}$ is the hydrostatic pressure.
849    
850     \subsubsection{Summary of equation sets supported by model}
851    
852     \paragraph{Atmosphere}
853    
854     Hydrostatic, and quasi-hydrostatic and quasi non-hydrostatic forms of the
855     compressible non-Boussinesq equations in $p-$coordinates are supported.
856    
857     \subparagraph{Hydrostatic and quasi-hydrostatic}
858    
859     The hydrostatic set is written out in $p-$coordinates in appendix Atmosphere
860     - see eq(\ref{eq:atmos-prime}).
861    
862     \subparagraph{Quasi-nonhydrostatic}
863    
864     A quasi-nonhydrostatic form is also supported.
865    
866     \paragraph{Ocean}
867    
868     \subparagraph{Hydrostatic and quasi-hydrostatic}
869    
870     Hydrostatic, and quasi-hydrostatic forms of the incompressible Boussinesq
871     equations in $z-$coordinates are supported.
872    
873     \subparagraph{Non-hydrostatic}
874    
875 adcroft 1.4 Non-hydrostatic forms of the incompressible Boussinesq equations in $z-$
876     coordinates are supported - see eqs(\ref{eq:ocean-mom}) to (\ref
877 cnh 1.1 {eq:ocean-salt}).
878    
879     \subsection{Solution strategy}
880    
881 adcroft 1.4 The method of solution employed in the \textbf{HPE}, \textbf{QH} and \textbf{
882 cnh 1.8 NH} models is summarized in Figure \ref{fig:solution-strategy}.
883     Under all dynamics, a 2-d elliptic equation is
884 cnh 1.1 first solved to find the surface pressure and the hydrostatic pressure at
885     any level computed from the weight of fluid above. Under \textbf{HPE} and
886     \textbf{QH} dynamics, the horizontal momentum equations are then stepped
887     forward and $\dot{r}$ found from continuity. Under \textbf{NH} dynamics a
888     3-d elliptic equation must be solved for the non-hydrostatic pressure before
889     stepping forward the horizontal momentum equations; $\dot{r}$ is found by
890     stepping forward the vertical momentum equation.
891    
892 cnh 1.3 %%CNHbegin
893 jmc 1.28 \input{s_overview/text/solution_strategy_figure.tex}
894 cnh 1.3 %%CNHend
895    
896 cnh 1.1 There is no penalty in implementing \textbf{QH} over \textbf{HPE} except, of
897 cnh 1.6 course, some complication that goes with the inclusion of $\cos \varphi \ $
898 cnh 1.1 Coriolis terms and the relaxation of the shallow atmosphere approximation.
899     But this leads to negligible increase in computation. In \textbf{NH}, in
900     contrast, one additional elliptic equation - a three-dimensional one - must
901     be inverted for $p_{nh}$. However the `overhead' of the \textbf{NH} model is
902     essentially negligible in the hydrostatic limit (see detailed discussion in
903     Marshall et al, 1997) resulting in a non-hydrostatic algorithm that, in the
904     hydrostatic limit, is as computationally economic as the \textbf{HPEs}.
905    
906     \subsection{Finding the pressure field}
907 cnh 1.7 \label{sec:finding_the_pressure_field}
908 cnh 1.1
909     Unlike the prognostic variables $u$, $v$, $w$, $\theta $ and $S$, the
910     pressure field must be obtained diagnostically. We proceed, as before, by
911     dividing the total (pressure/geo) potential in to three parts, a surface
912     part, $\phi _{s}(x,y)$, a hydrostatic part $\phi _{hyd}(x,y,r)$ and a
913     non-hydrostatic part $\phi _{nh}(x,y,r)$, as in (\ref{eq:phi-split}), and
914     writing the momentum equation as in (\ref{eq:mom-h}).
915    
916     \subsubsection{Hydrostatic pressure}
917    
918     Hydrostatic pressure is obtained by integrating (\ref{eq:hydrostatic})
919     vertically from $r=R_{o}$ where $\phi _{hyd}(r=R_{o})=0$, to yield:
920    
921     \begin{equation*}
922 adcroft 1.4 \int_{r}^{R_{o}}\frac{\partial \phi _{hyd}}{\partial r}dr=\left[ \phi _{hyd}
923 cnh 1.2 \right] _{r}^{R_{o}}=\int_{r}^{R_{o}}-bdr
924 cnh 1.1 \end{equation*}
925     and so
926    
927     \begin{equation}
928     \phi _{hyd}(x,y,r)=\int_{r}^{R_{o}}bdr \label{eq:hydro-phi}
929     \end{equation}
930    
931     The model can be easily modified to accommodate a loading term (e.g
932     atmospheric pressure pushing down on the ocean's surface) by setting:
933    
934     \begin{equation}
935     \phi _{hyd}(r=R_{o})=loading \label{eq:loading}
936     \end{equation}
937    
938     \subsubsection{Surface pressure}
939    
940 cnh 1.8 The surface pressure equation can be obtained by integrating continuity,
941     (\ref{eq:continuity}), vertically from $r=R_{fixed}$ to $r=R_{moving}$
942 cnh 1.1
943     \begin{equation*}
944 adcroft 1.4 \int_{R_{fixed}}^{R_{moving}}\left( \mathbf{\nabla }_{h}\cdot \vec{\mathbf{v}
945 cnh 1.2 }_{h}+\partial _{r}\dot{r}\right) dr=0
946 cnh 1.1 \end{equation*}
947    
948     Thus:
949    
950     \begin{equation*}
951     \frac{\partial \eta }{\partial t}+\vec{\mathbf{v}}.\nabla \eta
952 adcroft 1.4 +\int_{R_{fixed}}^{R_{moving}}\mathbf{\nabla }_{h}\cdot \vec{\mathbf{v}}
953 cnh 1.2 _{h}dr=0
954 cnh 1.1 \end{equation*}
955 adcroft 1.4 where $\eta =R_{moving}-R_{o}$ is the free-surface $r$-anomaly in units of $
956 cnh 1.1 r $. The above can be rearranged to yield, using Leibnitz's theorem:
957    
958     \begin{equation}
959     \frac{\partial \eta }{\partial t}+\mathbf{\nabla }_{h}\cdot
960     \int_{R_{fixed}}^{R_{moving}}\vec{\mathbf{v}}_{h}dr=\text{source}
961     \label{eq:free-surface}
962 adcroft 1.4 \end{equation}
963 cnh 1.1 where we have incorporated a source term.
964    
965     Whether $\phi $ is pressure (ocean model, $p/\rho _{c}$) or geopotential
966 cnh 1.8 (atmospheric model), in (\ref{eq:mom-h}), the horizontal gradient term can
967 cnh 1.1 be written
968     \begin{equation}
969 cnh 1.2 \mathbf{\nabla }_{h}\phi _{s}=\mathbf{\nabla }_{h}\left( b_{s}\eta \right)
970 cnh 1.1 \label{eq:phi-surf}
971 adcroft 1.4 \end{equation}
972 cnh 1.1 where $b_{s}$ is the buoyancy at the surface.
973    
974 cnh 1.8 In the hydrostatic limit ($\epsilon _{nh}=0$), equations (\ref{eq:mom-h}), (\ref
975 cnh 1.1 {eq:free-surface}) and (\ref{eq:phi-surf}) can be solved by inverting a 2-d
976     elliptic equation for $\phi _{s}$ as described in Chapter 2. Both `free
977     surface' and `rigid lid' approaches are available.
978    
979     \subsubsection{Non-hydrostatic pressure}
980    
981 cnh 1.8 Taking the horizontal divergence of (\ref{eq:mom-h}) and adding
982     $\frac{\partial }{\partial r}$ of (\ref{eq:mom-w}), invoking the continuity equation
983     (\ref{eq:continuity}), we deduce that:
984 cnh 1.1
985     \begin{equation}
986 adcroft 1.4 \nabla _{3}^{2}\phi _{nh}=\nabla .\vec{\mathbf{G}}_{\vec{v}}-\left( \mathbf{
987     \nabla }_{h}^{2}\phi _{s}+\mathbf{\nabla }^{2}\phi _{hyd}\right) =\nabla .
988 cnh 1.1 \vec{\mathbf{F}} \label{eq:3d-invert}
989     \end{equation}
990    
991     For a given rhs this 3-d elliptic equation must be inverted for $\phi _{nh}$
992     subject to appropriate choice of boundary conditions. This method is usually
993     called \textit{The Pressure Method} [Harlow and Welch, 1965; Williams, 1969;
994     Potter, 1976]. In the hydrostatic primitive equations case (\textbf{HPE}),
995     the 3-d problem does not need to be solved.
996    
997     \paragraph{Boundary Conditions}
998    
999     We apply the condition of no normal flow through all solid boundaries - the
1000     coasts (in the ocean) and the bottom:
1001    
1002     \begin{equation}
1003     \vec{\mathbf{v}}.\widehat{n}=0 \label{nonormalflow}
1004     \end{equation}
1005     where $\widehat{n}$ is a vector of unit length normal to the boundary. The
1006     kinematic condition (\ref{nonormalflow}) is also applied to the vertical
1007 adcroft 1.4 velocity at $r=R_{moving}$. No-slip $\left( v_{T}=0\right) \ $or slip $
1008 cnh 1.1 \left( \partial v_{T}/\partial n=0\right) \ $conditions are employed on the
1009     tangential component of velocity, $v_{T}$, at all solid boundaries,
1010     depending on the form chosen for the dissipative terms in the momentum
1011     equations - see below.
1012    
1013 cnh 1.8 Eq.(\ref{nonormalflow}) implies, making use of (\ref{eq:mom-h}), that:
1014 cnh 1.1
1015     \begin{equation}
1016     \widehat{n}.\nabla \phi _{nh}=\widehat{n}.\vec{\mathbf{F}}
1017     \label{eq:inhom-neumann-nh}
1018     \end{equation}
1019     where
1020    
1021     \begin{equation*}
1022     \vec{\mathbf{F}}=\vec{\mathbf{G}}_{\vec{v}}-\left( \mathbf{\nabla }_{h}\phi
1023     _{s}+\mathbf{\nabla }\phi _{hyd}\right)
1024 adcroft 1.4 \end{equation*}
1025 cnh 1.1 presenting inhomogeneous Neumann boundary conditions to the Elliptic problem
1026     (\ref{eq:3d-invert}). As shown, for example, by Williams (1969), one can
1027     exploit classical 3D potential theory and, by introducing an appropriately
1028 cnh 1.2 chosen $\delta $-function sheet of `source-charge', replace the
1029     inhomogeneous boundary condition on pressure by a homogeneous one. The
1030 adcroft 1.4 source term $rhs$ in (\ref{eq:3d-invert}) is the divergence of the vector $
1031     \vec{\mathbf{F}}.$ By simultaneously setting $
1032 cnh 1.1 \begin{array}{l}
1033 adcroft 1.4 \widehat{n}.\vec{\mathbf{F}}
1034     \end{array}
1035 cnh 1.1 =0$\ and $\widehat{n}.\nabla \phi _{nh}=0\ $on the boundary the following
1036 cnh 1.2 self-consistent but simpler homogenized Elliptic problem is obtained:
1037 cnh 1.1
1038     \begin{equation*}
1039 cnh 1.2 \nabla ^{2}\phi _{nh}=\nabla .\widetilde{\vec{\mathbf{F}}}\qquad
1040 adcroft 1.4 \end{equation*}
1041 cnh 1.1 where $\widetilde{\vec{\mathbf{F}}}$ is a modified $\vec{\mathbf{F}}$ such
1042 adcroft 1.4 that $\widetilde{\vec{\mathbf{F}}}.\widehat{n}=0$. As is implied by (\ref
1043 cnh 1.1 {eq:inhom-neumann-nh}) the modified boundary condition becomes:
1044    
1045     \begin{equation}
1046     \widehat{n}.\nabla \phi _{nh}=0 \label{eq:hom-neumann-nh}
1047     \end{equation}
1048    
1049     If the flow is `close' to hydrostatic balance then the 3-d inversion
1050     converges rapidly because $\phi _{nh}\ $is then only a small correction to
1051     the hydrostatic pressure field (see the discussion in Marshall et al, a,b).
1052    
1053 cnh 1.8 The solution $\phi _{nh}\ $to (\ref{eq:3d-invert}) and (\ref{eq:inhom-neumann-nh})
1054 cnh 1.1 does not vanish at $r=R_{moving}$, and so refines the pressure there.
1055    
1056     \subsection{Forcing/dissipation}
1057    
1058     \subsubsection{Forcing}
1059    
1060     The forcing terms $\mathcal{F}$ on the rhs of the equations are provided by
1061 cnh 1.8 `physics packages' and forcing packages. These are described later on.
1062 cnh 1.1
1063     \subsubsection{Dissipation}
1064    
1065     \paragraph{Momentum}
1066    
1067     Many forms of momentum dissipation are available in the model. Laplacian and
1068     biharmonic frictions are commonly used:
1069    
1070     \begin{equation}
1071 adcroft 1.4 D_{V}=A_{h}\nabla _{h}^{2}v+A_{v}\frac{\partial ^{2}v}{\partial z^{2}}
1072 cnh 1.1 +A_{4}\nabla _{h}^{4}v \label{eq:dissipation}
1073     \end{equation}
1074     where $A_{h}$ and $A_{v}\ $are (constant) horizontal and vertical viscosity
1075     coefficients and $A_{4}\ $is the horizontal coefficient for biharmonic
1076     friction. These coefficients are the same for all velocity components.
1077    
1078     \paragraph{Tracers}
1079    
1080     The mixing terms for the temperature and salinity equations have a similar
1081     form to that of momentum except that the diffusion tensor can be
1082 edhill 1.26 non-diagonal and have varying coefficients.
1083 cnh 1.1 \begin{equation}
1084     D_{T,S}=\nabla .[\underline{\underline{K}}\nabla (T,S)]+K_{4}\nabla
1085     _{h}^{4}(T,S) \label{eq:diffusion}
1086     \end{equation}
1087 adcroft 1.4 where $\underline{\underline{K}}\ $is the diffusion tensor and the $K_{4}\ $
1088 cnh 1.1 horizontal coefficient for biharmonic diffusion. In the simplest case where
1089     the subgrid-scale fluxes of heat and salt are parameterized with constant
1090     horizontal and vertical diffusion coefficients, $\underline{\underline{K}}$,
1091     reduces to a diagonal matrix with constant coefficients:
1092    
1093     \begin{equation}
1094     \qquad \qquad \qquad \qquad K=\left(
1095     \begin{array}{ccc}
1096     K_{h} & 0 & 0 \\
1097     0 & K_{h} & 0 \\
1098 adcroft 1.4 0 & 0 & K_{v}
1099 cnh 1.1 \end{array}
1100     \right) \qquad \qquad \qquad \label{eq:diagonal-diffusion-tensor}
1101     \end{equation}
1102     where $K_{h}\ $and $K_{v}\ $are the horizontal and vertical diffusion
1103     coefficients. These coefficients are the same for all tracers (temperature,
1104     salinity ... ).
1105    
1106     \subsection{Vector invariant form}
1107    
1108 edhill 1.21 For some purposes it is advantageous to write momentum advection in
1109     eq(\ref {eq:horizontal_mtm}) and (\ref{eq:vertical_mtm}) in the
1110     (so-called) `vector invariant' form:
1111 cnh 1.1
1112     \begin{equation}
1113 adcroft 1.4 \frac{D\vec{\mathbf{v}}}{Dt}=\frac{\partial \vec{\mathbf{v}}}{\partial t}
1114     +\left( \nabla \times \vec{\mathbf{v}}\right) \times \vec{\mathbf{v}}+\nabla
1115 cnh 1.2 \left[ \frac{1}{2}(\vec{\mathbf{v}}\cdot \vec{\mathbf{v}})\right]
1116 cnh 1.1 \label{eq:vi-identity}
1117 adcroft 1.4 \end{equation}
1118 cnh 1.1 This permits alternative numerical treatments of the non-linear terms based
1119     on their representation as a vorticity flux. Because gradients of coordinate
1120     vectors no longer appear on the rhs of (\ref{eq:vi-identity}), explicit
1121 adcroft 1.4 representation of the metric terms in (\ref{eq:gu-speherical}), (\ref
1122 cnh 1.1 {eq:gv-spherical}) and (\ref{eq:gw-spherical}), can be avoided: information
1123     about the geometry is contained in the areas and lengths of the volumes used
1124     to discretize the model.
1125    
1126     \subsection{Adjoint}
1127    
1128 cnh 1.8 Tangent linear and adjoint counterparts of the forward model are described
1129 cnh 1.2 in Chapter 5.
1130 cnh 1.1
1131 jmc 1.28 % $Header: /u/gcmpack/manual/s_overview/text/manual.tex,v 1.27 2008/01/17 21:28:22 jmc Exp $
1132 cnh 1.1 % $Name: $
1133    
1134     \section{Appendix ATMOSPHERE}
1135    
1136     \subsection{Hydrostatic Primitive Equations for the Atmosphere in pressure
1137     coordinates}
1138    
1139     \label{sect-hpe-p}
1140    
1141     The hydrostatic primitive equations (HPEs) in p-coordinates are:
1142     \begin{eqnarray}
1143 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1144 cnh 1.2 _{h}+\mathbf{\nabla }_{p}\phi &=&\vec{\mathbf{\mathcal{F}}}
1145 cnh 1.1 \label{eq:atmos-mom} \\
1146 cnh 1.2 \frac{\partial \phi }{\partial p}+\alpha &=&0 \label{eq-p-hydro-start} \\
1147 adcroft 1.4 \mathbf{\nabla }_{p}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \omega }{
1148 cnh 1.1 \partial p} &=&0 \label{eq:atmos-cont} \\
1149 cnh 1.2 p\alpha &=&RT \label{eq:atmos-eos} \\
1150 cnh 1.1 c_{v}\frac{DT}{Dt}+p\frac{D\alpha }{Dt} &=&\mathcal{Q} \label{eq:atmos-heat}
1151 adcroft 1.4 \end{eqnarray}
1152 cnh 1.1 where $\vec{\mathbf{v}}_{h}=(u,v,0)$ is the `horizontal' (on pressure
1153     surfaces) component of velocity,$\frac{D}{Dt}=\vec{\mathbf{v}}_{h}\cdot
1154     \mathbf{\nabla }_{p}+\omega \frac{\partial }{\partial p}$ is the total
1155 cnh 1.6 derivative, $f=2\Omega \sin \varphi$ is the Coriolis parameter, $\phi =gz$ is
1156 adcroft 1.4 the geopotential, $\alpha =1/\rho $ is the specific volume, $\omega =\frac{Dp
1157     }{Dt}$ is the vertical velocity in the $p-$coordinate. Equation(\ref
1158     {eq:atmos-heat}) is the first law of thermodynamics where internal energy $
1159     e=c_{v}T$, $T$ is temperature, $Q$ is the rate of heating per unit mass and $
1160 cnh 1.1 p\frac{D\alpha }{Dt}$ is the work done by the fluid in compressing.
1161    
1162     It is convenient to cast the heat equation in terms of potential temperature
1163     $\theta $ so that it looks more like a generic conservation law.
1164     Differentiating (\ref{eq:atmos-eos}) we get:
1165     \begin{equation*}
1166     p\frac{D\alpha }{Dt}+\alpha \frac{Dp}{Dt}=R\frac{DT}{Dt}
1167 adcroft 1.4 \end{equation*}
1168     which, when added to the heat equation (\ref{eq:atmos-heat}) and using $
1169 cnh 1.1 c_{p}=c_{v}+R$, gives:
1170     \begin{equation}
1171     c_{p}\frac{DT}{Dt}-\alpha \frac{Dp}{Dt}=\mathcal{Q}
1172     \label{eq-p-heat-interim}
1173 adcroft 1.4 \end{equation}
1174 cnh 1.1 Potential temperature is defined:
1175     \begin{equation}
1176     \theta =T(\frac{p_{c}}{p})^{\kappa } \label{eq:potential-temp}
1177 adcroft 1.4 \end{equation}
1178 cnh 1.1 where $p_{c}$ is a reference pressure and $\kappa =R/c_{p}$. For convenience
1179     we will make use of the Exner function $\Pi (p)$ which defined by:
1180     \begin{equation}
1181     \Pi (p)=c_{p}(\frac{p}{p_{c}})^{\kappa } \label{Exner}
1182 adcroft 1.4 \end{equation}
1183 cnh 1.1 The following relations will be useful and are easily expressed in terms of
1184     the Exner function:
1185     \begin{equation*}
1186     c_{p}T=\Pi \theta \;\;;\;\;\frac{\partial \Pi }{\partial p}=\frac{\kappa \Pi
1187 adcroft 1.4 }{p}\;\;;\;\;\alpha =\frac{\kappa \Pi \theta }{p}=\frac{\partial \ \Pi }{
1188     \partial p}\theta \;\;;\;\;\frac{D\Pi }{Dt}=\frac{\partial \Pi }{\partial p}
1189 cnh 1.1 \frac{Dp}{Dt}
1190 adcroft 1.4 \end{equation*}
1191 cnh 1.1 where $b=\frac{\partial \ \Pi }{\partial p}\theta $ is the buoyancy.
1192    
1193     The heat equation is obtained by noting that
1194     \begin{equation*}
1195     c_{p}\frac{DT}{Dt}=\frac{D(\Pi \theta )}{Dt}=\Pi \frac{D\theta }{Dt}+\theta
1196 cnh 1.2 \frac{D\Pi }{Dt}=\Pi \frac{D\theta }{Dt}+\alpha \frac{Dp}{Dt}
1197 cnh 1.1 \end{equation*}
1198     and on substituting into (\ref{eq-p-heat-interim}) gives:
1199     \begin{equation}
1200     \Pi \frac{D\theta }{Dt}=\mathcal{Q}
1201     \label{eq:potential-temperature-equation}
1202     \end{equation}
1203     which is in conservative form.
1204    
1205 adcroft 1.4 For convenience in the model we prefer to step forward (\ref
1206 cnh 1.1 {eq:potential-temperature-equation}) rather than (\ref{eq:atmos-heat}).
1207    
1208     \subsubsection{Boundary conditions}
1209    
1210     The upper and lower boundary conditions are :
1211     \begin{eqnarray}
1212     \mbox{at the top:}\;\;p=0 &&\text{, }\omega =\frac{Dp}{Dt}=0 \\
1213     \mbox{at the surface:}\;\;p=p_{s} &&\text{, }\phi =\phi _{topo}=g~Z_{topo}
1214     \label{eq:boundary-condition-atmosphere}
1215     \end{eqnarray}
1216     In $p$-coordinates, the upper boundary acts like a solid boundary ($\omega
1217     =0 $); in $z$-coordinates and the lower boundary is analogous to a free
1218     surface ($\phi $ is imposed and $\omega \neq 0$).
1219    
1220     \subsubsection{Splitting the geo-potential}
1221 jmc 1.22 \label{sec:hpe-p-geo-potential-split}
1222 cnh 1.1
1223     For the purposes of initialization and reducing round-off errors, the model
1224     deals with perturbations from reference (or ``standard'') profiles. For
1225     example, the hydrostatic geopotential associated with the resting atmosphere
1226     is not dynamically relevant and can therefore be subtracted from the
1227     equations. The equations written in terms of perturbations are obtained by
1228     substituting the following definitions into the previous model equations:
1229     \begin{eqnarray}
1230     \theta &=&\theta _{o}+\theta ^{\prime } \label{eq:atmos-ref-prof-theta} \\
1231     \alpha &=&\alpha _{o}+\alpha ^{\prime } \label{eq:atmos-ref-prof-alpha} \\
1232     \phi &=&\phi _{o}+\phi ^{\prime } \label{eq:atmos-ref-prof-phi}
1233     \end{eqnarray}
1234     The reference state (indicated by subscript ``0'') corresponds to
1235     horizontally homogeneous atmosphere at rest ($\theta _{o},\alpha _{o},\phi
1236     _{o}$) with surface pressure $p_{o}(x,y)$ that satisfies $\phi
1237     _{o}(p_{o})=g~Z_{topo}$, defined:
1238     \begin{eqnarray*}
1239     \theta _{o}(p) &=&f^{n}(p) \\
1240     \alpha _{o}(p) &=&\Pi _{p}\theta _{o} \\
1241     \phi _{o}(p) &=&\phi _{topo}-\int_{p_{0}}^{p}\alpha _{o}dp
1242     \end{eqnarray*}
1243     %\begin{eqnarray*}
1244     %\phi'_\alpha & = & \int^p_{p_o} (\alpha_o -\alpha) dp \\
1245     %\phi'_s(x,y,t) & = & \int_{p_o}^{p_s} \alpha dp
1246     %\end{eqnarray*}
1247    
1248     The final form of the HPE's in p coordinates is then:
1249     \begin{eqnarray}
1250 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1251 edhill 1.21 _{h}+\mathbf{\nabla }_{p}\phi ^{\prime } &=&\vec{\mathbf{\mathcal{F}}}
1252     \label{eq:atmos-prime} \\
1253 cnh 1.1 \frac{\partial \phi ^{\prime }}{\partial p}+\alpha ^{\prime } &=&0 \\
1254 adcroft 1.4 \mathbf{\nabla }_{p}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \omega }{
1255 cnh 1.1 \partial p} &=&0 \\
1256     \frac{\partial \Pi }{\partial p}\theta ^{\prime } &=&\alpha ^{\prime } \\
1257 cnh 1.8 \frac{D\theta }{Dt} &=&\frac{\mathcal{Q}}{\Pi }
1258 cnh 1.1 \end{eqnarray}
1259    
1260 jmc 1.28 % $Header: /u/gcmpack/manual/s_overview/text/manual.tex,v 1.27 2008/01/17 21:28:22 jmc Exp $
1261 cnh 1.1 % $Name: $
1262    
1263     \section{Appendix OCEAN}
1264    
1265     \subsection{Equations of motion for the ocean}
1266    
1267     We review here the method by which the standard (Boussinesq, incompressible)
1268     HPE's for the ocean written in z-coordinates are obtained. The
1269     non-Boussinesq equations for oceanic motion are:
1270     \begin{eqnarray}
1271 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1272 cnh 1.1 _{h}+\frac{1}{\rho }\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}} \\
1273     \epsilon _{nh}\frac{Dw}{Dt}+g+\frac{1}{\rho }\frac{\partial p}{\partial z}
1274     &=&\epsilon _{nh}\mathcal{F}_{w} \\
1275 adcroft 1.4 \frac{1}{\rho }\frac{D\rho }{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}
1276 cnh 1.8 _{h}+\frac{\partial w}{\partial z} &=&0 \label{eq-zns-cont}\\
1277     \rho &=&\rho (\theta ,S,p) \label{eq-zns-eos}\\
1278     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-zns-heat}\\
1279     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-zns-salt}
1280     \label{eq:non-boussinesq}
1281 adcroft 1.4 \end{eqnarray}
1282 cnh 1.1 These equations permit acoustics modes, inertia-gravity waves,
1283 cnh 1.10 non-hydrostatic motions, a geostrophic (Rossby) mode and a thermohaline
1284 cnh 1.1 mode. As written, they cannot be integrated forward consistently - if we
1285     step $\rho $ forward in (\ref{eq-zns-cont}), the answer will not be
1286 adcroft 1.4 consistent with that obtained by stepping (\ref{eq-zns-heat}) and (\ref
1287 cnh 1.1 {eq-zns-salt}) and then using (\ref{eq-zns-eos}) to yield $\rho $. It is
1288     therefore necessary to manipulate the system as follows. Differentiating the
1289     EOS (equation of state) gives:
1290    
1291     \begin{equation}
1292     \frac{D\rho }{Dt}=\left. \frac{\partial \rho }{\partial \theta }\right|
1293     _{S,p}\frac{D\theta }{Dt}+\left. \frac{\partial \rho }{\partial S}\right|
1294     _{\theta ,p}\frac{DS}{Dt}+\left. \frac{\partial \rho }{\partial p}\right|
1295     _{\theta ,S}\frac{Dp}{Dt} \label{EOSexpansion}
1296     \end{equation}
1297    
1298 edhill 1.21 Note that $\frac{\partial \rho }{\partial p}=\frac{1}{c_{s}^{2}}$ is
1299     the reciprocal of the sound speed ($c_{s}$) squared. Substituting into
1300     \ref{eq-zns-cont} gives:
1301 cnh 1.1 \begin{equation}
1302 adcroft 1.4 \frac{1}{\rho c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{
1303 cnh 1.1 v}}+\partial _{z}w\approx 0 \label{eq-zns-pressure}
1304     \end{equation}
1305     where we have used an approximation sign to indicate that we have assumed
1306     adiabatic motion, dropping the $\frac{D\theta }{Dt}$ and $\frac{DS}{Dt}$.
1307     Replacing \ref{eq-zns-cont} with \ref{eq-zns-pressure} yields a system that
1308     can be explicitly integrated forward:
1309     \begin{eqnarray}
1310 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1311 cnh 1.1 _{h}+\frac{1}{\rho }\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1312     \label{eq-cns-hmom} \\
1313     \epsilon _{nh}\frac{Dw}{Dt}+g+\frac{1}{\rho }\frac{\partial p}{\partial z}
1314     &=&\epsilon _{nh}\mathcal{F}_{w} \label{eq-cns-hydro} \\
1315 adcroft 1.4 \frac{1}{\rho c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{
1316 cnh 1.1 v}}_{h}+\frac{\partial w}{\partial z} &=&0 \label{eq-cns-cont} \\
1317     \rho &=&\rho (\theta ,S,p) \label{eq-cns-eos} \\
1318     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-cns-heat} \\
1319     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-cns-salt}
1320     \end{eqnarray}
1321    
1322     \subsubsection{Compressible z-coordinate equations}
1323    
1324     Here we linearize the acoustic modes by replacing $\rho $ with $\rho _{o}(z)$
1325     wherever it appears in a product (ie. non-linear term) - this is the
1326     `Boussinesq assumption'. The only term that then retains the full variation
1327     in $\rho $ is the gravitational acceleration:
1328     \begin{eqnarray}
1329 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1330 cnh 1.1 _{h}+\frac{1}{\rho _{o}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1331     \label{eq-zcb-hmom} \\
1332 adcroft 1.4 \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{o}}+\frac{1}{\rho _{o}}
1333 cnh 1.1 \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1334     \label{eq-zcb-hydro} \\
1335 adcroft 1.4 \frac{1}{\rho _{o}c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{
1336 cnh 1.1 \mathbf{v}}_{h}+\frac{\partial w}{\partial z} &=&0 \label{eq-zcb-cont} \\
1337     \rho &=&\rho (\theta ,S,p) \label{eq-zcb-eos} \\
1338     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-zcb-heat} \\
1339     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-zcb-salt}
1340     \end{eqnarray}
1341     These equations still retain acoustic modes. But, because the
1342 adcroft 1.4 ``compressible'' terms are linearized, the pressure equation \ref
1343 cnh 1.1 {eq-zcb-cont} can be integrated implicitly with ease (the time-dependent
1344     term appears as a Helmholtz term in the non-hydrostatic pressure equation).
1345     These are the \emph{truly} compressible Boussinesq equations. Note that the
1346     EOS must have the same pressure dependency as the linearized pressure term,
1347 adcroft 1.4 ie. $\left. \frac{\partial \rho }{\partial p}\right| _{\theta ,S}=\frac{1}{
1348 cnh 1.1 c_{s}^{2}}$, for consistency.
1349    
1350     \subsubsection{`Anelastic' z-coordinate equations}
1351    
1352     The anelastic approximation filters the acoustic mode by removing the
1353 adcroft 1.4 time-dependency in the continuity (now pressure-) equation (\ref{eq-zcb-cont}
1354     ). This could be done simply by noting that $\frac{Dp}{Dt}\approx -g\rho _{o}
1355 cnh 1.1 \frac{Dz}{Dt}=-g\rho _{o}w$, but this leads to an inconsistency between
1356     continuity and EOS. A better solution is to change the dependency on
1357     pressure in the EOS by splitting the pressure into a reference function of
1358     height and a perturbation:
1359     \begin{equation*}
1360 cnh 1.2 \rho =\rho (\theta ,S,p_{o}(z)+\epsilon _{s}p^{\prime })
1361 cnh 1.1 \end{equation*}
1362     Remembering that the term $\frac{Dp}{Dt}$ in continuity comes from
1363     differentiating the EOS, the continuity equation then becomes:
1364     \begin{equation*}
1365 adcroft 1.4 \frac{1}{\rho _{o}c_{s}^{2}}\left( \frac{Dp_{o}}{Dt}+\epsilon _{s}\frac{
1366     Dp^{\prime }}{Dt}\right) +\mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+
1367 cnh 1.2 \frac{\partial w}{\partial z}=0
1368 cnh 1.1 \end{equation*}
1369     If the time- and space-scales of the motions of interest are longer than
1370 adcroft 1.4 those of acoustic modes, then $\frac{Dp^{\prime }}{Dt}<<(\frac{Dp_{o}}{Dt},
1371 cnh 1.1 \mathbf{\nabla }\cdot \vec{\mathbf{v}}_{h})$ in the continuity equations and
1372 adcroft 1.4 $\left. \frac{\partial \rho }{\partial p}\right| _{\theta ,S}\frac{
1373 cnh 1.1 Dp^{\prime }}{Dt}<<\left. \frac{\partial \rho }{\partial p}\right| _{\theta
1374     ,S}\frac{Dp_{o}}{Dt}$ in the EOS (\ref{EOSexpansion}). Thus we set $\epsilon
1375     _{s}=0$, removing the dependency on $p^{\prime }$ in the continuity equation
1376     and EOS. Expanding $\frac{Dp_{o}(z)}{Dt}=-g\rho _{o}w$ then leads to the
1377     anelastic continuity equation:
1378     \begin{equation}
1379 adcroft 1.4 \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z}-
1380 cnh 1.1 \frac{g}{c_{s}^{2}}w=0 \label{eq-za-cont1}
1381     \end{equation}
1382     A slightly different route leads to the quasi-Boussinesq continuity equation
1383 adcroft 1.4 where we use the scaling $\frac{\partial \rho ^{\prime }}{\partial t}+
1384     \mathbf{\nabla }_{3}\cdot \rho ^{\prime }\vec{\mathbf{v}}<<\mathbf{\nabla }
1385 cnh 1.1 _{3}\cdot \rho _{o}\vec{\mathbf{v}}$ yielding:
1386     \begin{equation}
1387 adcroft 1.4 \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{1}{\rho _{o}}\frac{
1388 cnh 1.1 \partial \left( \rho _{o}w\right) }{\partial z}=0 \label{eq-za-cont2}
1389     \end{equation}
1390     Equations \ref{eq-za-cont1} and \ref{eq-za-cont2} are in fact the same
1391     equation if:
1392     \begin{equation}
1393     \frac{1}{\rho _{o}}\frac{\partial \rho _{o}}{\partial z}=\frac{-g}{c_{s}^{2}}
1394     \end{equation}
1395     Again, note that if $\rho _{o}$ is evaluated from prescribed $\theta _{o}$
1396 adcroft 1.4 and $S_{o}$ profiles, then the EOS dependency on $p_{o}$ and the term $\frac{
1397 cnh 1.1 g}{c_{s}^{2}}$ in continuity should be referred to those same profiles. The
1398     full set of `quasi-Boussinesq' or `anelastic' equations for the ocean are
1399     then:
1400     \begin{eqnarray}
1401 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1402 cnh 1.1 _{h}+\frac{1}{\rho _{o}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1403     \label{eq-zab-hmom} \\
1404 adcroft 1.4 \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{o}}+\frac{1}{\rho _{o}}
1405 cnh 1.1 \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1406     \label{eq-zab-hydro} \\
1407 adcroft 1.4 \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{1}{\rho _{o}}\frac{
1408 cnh 1.1 \partial \left( \rho _{o}w\right) }{\partial z} &=&0 \label{eq-zab-cont} \\
1409     \rho &=&\rho (\theta ,S,p_{o}(z)) \label{eq-zab-eos} \\
1410     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-zab-heat} \\
1411     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-zab-salt}
1412     \end{eqnarray}
1413    
1414     \subsubsection{Incompressible z-coordinate equations}
1415    
1416     Here, the objective is to drop the depth dependence of $\rho _{o}$ and so,
1417     technically, to also remove the dependence of $\rho $ on $p_{o}$. This would
1418     yield the ``truly'' incompressible Boussinesq equations:
1419     \begin{eqnarray}
1420 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1421 cnh 1.1 _{h}+\frac{1}{\rho _{c}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1422     \label{eq-ztb-hmom} \\
1423 adcroft 1.4 \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{c}}+\frac{1}{\rho _{c}}
1424 cnh 1.1 \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1425     \label{eq-ztb-hydro} \\
1426     \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z}
1427     &=&0 \label{eq-ztb-cont} \\
1428     \rho &=&\rho (\theta ,S) \label{eq-ztb-eos} \\
1429     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-ztb-heat} \\
1430     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-ztb-salt}
1431     \end{eqnarray}
1432     where $\rho _{c}$ is a constant reference density of water.
1433    
1434     \subsubsection{Compressible non-divergent equations}
1435    
1436     The above ``incompressible'' equations are incompressible in both the flow
1437     and the density. In many oceanic applications, however, it is important to
1438     retain compressibility effects in the density. To do this we must split the
1439     density thus:
1440     \begin{equation*}
1441     \rho =\rho _{o}+\rho ^{\prime }
1442 adcroft 1.4 \end{equation*}
1443 cnh 1.1 We then assert that variations with depth of $\rho _{o}$ are unimportant
1444     while the compressible effects in $\rho ^{\prime }$ are:
1445     \begin{equation*}
1446     \rho _{o}=\rho _{c}
1447 adcroft 1.4 \end{equation*}
1448 cnh 1.1 \begin{equation*}
1449     \rho ^{\prime }=\rho (\theta ,S,p_{o}(z))-\rho _{o}
1450 adcroft 1.4 \end{equation*}
1451 cnh 1.1 This then yields what we can call the semi-compressible Boussinesq
1452     equations:
1453     \begin{eqnarray}
1454 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1455     _{h}+\frac{1}{\rho _{c}}\mathbf{\nabla }_{z}p^{\prime } &=&\vec{\mathbf{
1456 cnh 1.1 \mathcal{F}}} \label{eq:ocean-mom} \\
1457     \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho ^{\prime }}{\rho _{c}}+\frac{1}{\rho
1458     _{c}}\frac{\partial p^{\prime }}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1459     \label{eq:ocean-wmom} \\
1460     \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z}
1461     &=&0 \label{eq:ocean-cont} \\
1462     \rho ^{\prime } &=&\rho (\theta ,S,p_{o}(z))-\rho _{c} \label{eq:ocean-eos}
1463     \\
1464     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq:ocean-theta} \\
1465     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq:ocean-salt}
1466 adcroft 1.4 \end{eqnarray}
1467 cnh 1.1 Note that the hydrostatic pressure of the resting fluid, including that
1468     associated with $\rho _{c}$, is subtracted out since it has no effect on the
1469     dynamics.
1470    
1471     Though necessary, the assumptions that go into these equations are messy
1472     since we essentially assume a different EOS for the reference density and
1473     the perturbation density. Nevertheless, it is the hydrostatic ($\epsilon
1474     _{nh}=0$ form of these equations that are used throughout the ocean modeling
1475     community and referred to as the primitive equations (HPE).
1476    
1477 jmc 1.28 % $Header: /u/gcmpack/manual/s_overview/text/manual.tex,v 1.27 2008/01/17 21:28:22 jmc Exp $
1478 cnh 1.1 % $Name: $
1479    
1480     \section{Appendix:OPERATORS}
1481    
1482     \subsection{Coordinate systems}
1483    
1484     \subsubsection{Spherical coordinates}
1485    
1486     In spherical coordinates, the velocity components in the zonal, meridional
1487     and vertical direction respectively, are given by (see Fig.2) :
1488    
1489     \begin{equation*}
1490 cnh 1.6 u=r\cos \varphi \frac{D\lambda }{Dt}
1491 cnh 1.1 \end{equation*}
1492    
1493     \begin{equation*}
1494 edhill 1.26 v=r\frac{D\varphi }{Dt}
1495 cnh 1.1 \end{equation*}
1496    
1497     \begin{equation*}
1498 cnh 1.2 \dot{r}=\frac{Dr}{Dt}
1499 cnh 1.1 \end{equation*}
1500    
1501 cnh 1.6 Here $\varphi $ is the latitude, $\lambda $ the longitude, $r$ the radial
1502 cnh 1.1 distance of the particle from the center of the earth, $\Omega $ is the
1503     angular speed of rotation of the Earth and $D/Dt$ is the total derivative.
1504    
1505 edhill 1.26 The `grad' ($\nabla $) and `div' ($\nabla\cdot$) operators are defined by, in
1506 cnh 1.1 spherical coordinates:
1507    
1508     \begin{equation*}
1509 cnh 1.6 \nabla \equiv \left( \frac{1}{r\cos \varphi }\frac{\partial }{\partial \lambda }
1510     ,\frac{1}{r}\frac{\partial }{\partial \varphi },\frac{\partial }{\partial r}
1511 cnh 1.2 \right)
1512 cnh 1.1 \end{equation*}
1513    
1514     \begin{equation*}
1515 edhill 1.26 \nabla\cdot v\equiv \frac{1}{r\cos \varphi }\left\{ \frac{\partial u}{\partial
1516 cnh 1.6 \lambda }+\frac{\partial }{\partial \varphi }\left( v\cos \varphi \right) \right\}
1517 cnh 1.2 +\frac{1}{r^{2}}\frac{\partial \left( r^{2}\dot{r}\right) }{\partial r}
1518 cnh 1.1 \end{equation*}
1519    
1520 adcroft 1.4 %tci%\end{document}

  ViewVC Help
Powered by ViewVC 1.1.22