/[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 1.30 - (hide annotations) (download) (as text)
Wed May 11 18:45:43 2016 UTC (9 years, 2 months ago) by jmc
Branch: MAIN
CVS Tags: checkpoint01, HEAD
Changes since 1.29: +9 -27 lines
File MIME type: application/x-tex
fix definition of total derivative in ATMOSPHERE appendix (section 1.4)

1 jmc 1.30 % $Header: /u/gcmpack/manual/s_overview/text/manual.tex,v 1.29 2010/08/30 23:09:21 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 cnh 1.16 This document provides the reader with the information necessary to
38 cnh 1.1 carry out numerical experiments using MITgcm. It gives a comprehensive
39     description of the continuous equations on which the model is based, the
40     numerical algorithms the model employs and a description of the associated
41     program code. Along with the hydrodynamical kernel, physical and
42     biogeochemical parameterizations of key atmospheric and oceanic processes
43     are available. A number of examples illustrating the use of the model in
44     both process and general circulation studies of the atmosphere and ocean are
45     also presented.
46    
47 cnh 1.16 \section{Introduction}
48 afe 1.18 \begin{rawhtml}
49 afe 1.19 <!-- CMIREDIR:innovations: -->
50 afe 1.18 \end{rawhtml}
51    
52 cnh 1.16
53 cnh 1.1 MITgcm has a number of novel aspects:
54    
55     \begin{itemize}
56     \item it can be used to study both atmospheric and oceanic phenomena; one
57     hydrodynamical kernel is used to drive forward both atmospheric and oceanic
58 cnh 1.7 models - see fig \ref{fig:onemodel}
59 cnh 1.1
60 cnh 1.3 %% CNHbegin
61 jmc 1.28 \input{s_overview/text/one_model_figure}
62 cnh 1.3 %% CNHend
63    
64 cnh 1.1 \item it has a non-hydrostatic capability and so can be used to study both
65 cnh 1.7 small-scale and large scale processes - see fig \ref{fig:all-scales}
66 cnh 1.1
67 cnh 1.3 %% CNHbegin
68 jmc 1.28 \input{s_overview/text/all_scales_figure}
69 cnh 1.3 %% CNHend
70    
71 cnh 1.1 \item finite volume techniques are employed yielding an intuitive
72     discretization and support for the treatment of irregular geometries using
73 cnh 1.7 orthogonal curvilinear grids and shaved cells - see fig \ref{fig:finite-volumes}
74 cnh 1.3
75     %% CNHbegin
76 jmc 1.28 \input{s_overview/text/fvol_figure}
77 cnh 1.3 %% CNHend
78 cnh 1.1
79     \item tangent linear and adjoint counterparts are automatically maintained
80     along with the forward model, permitting sensitivity and optimization
81     studies.
82    
83     \item the model is developed to perform efficiently on a wide variety of
84     computational platforms.
85     \end{itemize}
86    
87 jmc 1.27
88 cnh 1.16 Key publications reporting on and charting the development of the model are
89 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}
90     (an overview on the model formulation can also be found in \cite{adcroft:04c}):
91 cnh 1.12
92     \begin{verbatim}
93     Hill, C. and J. Marshall, (1995)
94     Application of a Parallel Navier-Stokes Model to Ocean Circulation in
95     Parallel Computational Fluid Dynamics
96     In Proceedings of Parallel Computational Fluid Dynamics: Implementations
97     and Results Using Parallel Computers, 545-552.
98     Elsevier Science B.V.: New York
99    
100     Marshall, J., C. Hill, L. Perelman, and A. Adcroft, (1997)
101 cnh 1.16 Hydrostatic, quasi-hydrostatic, and nonhydrostatic ocean modeling
102 cnh 1.12 J. Geophysical Res., 102(C3), 5733-5752.
103    
104     Marshall, J., A. Adcroft, C. Hill, L. Perelman, and C. Heisey, (1997)
105     A finite-volume, incompressible Navier Stokes model for studies of the ocean
106     on parallel computers,
107     J. Geophysical Res., 102(C3), 5753-5766.
108    
109     Adcroft, A.J., Hill, C.N. and J. Marshall, (1997)
110     Representation of topography by shaved cells in a height coordinate ocean
111     model
112     Mon Wea Rev, vol 125, 2293-2315
113    
114     Marshall, J., Jones, H. and C. Hill, (1998)
115     Efficient ocean modeling using non-hydrostatic algorithms
116     Journal of Marine Systems, 18, 115-134
117    
118     Adcroft, A., Hill C. and J. Marshall: (1999)
119     A new treatment of the Coriolis terms in C-grid models at both high and low
120     resolutions,
121     Mon. Wea. Rev. Vol 127, pages 1928-1936
122    
123     Hill, C, Adcroft,A., Jamous,D., and J. Marshall, (1999)
124     A Strategy for Terascale Climate Modeling.
125 cnh 1.14 In Proceedings of the Eighth ECMWF Workshop on the Use of Parallel Processors
126     in Meteorology, pages 406-425
127     World Scientific Publishing Co: UK
128 cnh 1.12
129     Marotzke, J, Giering,R., Zhang, K.Q., Stammer,D., Hill,C., and T.Lee, (1999)
130     Construction of the adjoint MIT ocean general circulation model and
131     application to Atlantic heat transport variability
132     J. Geophysical Res., 104(C12), 29,529-29,547.
133    
134     \end{verbatim}
135 cnh 1.1
136     We begin by briefly showing some of the results of the model in action to
137     give a feel for the wide range of problems that can be addressed using it.
138    
139     \section{Illustrations of the model in action}
140    
141 edhill 1.24 MITgcm has been designed and used to model a wide range of phenomena,
142 cnh 1.1 from convection on the scale of meters in the ocean to the global pattern of
143 cnh 1.7 atmospheric winds - see figure \ref{fig:all-scales}. To give a flavor of the
144 cnh 1.1 kinds of problems the model has been used to study, we briefly describe some
145     of them here. A more detailed description of the underlying formulation,
146     numerical algorithm and implementation that lie behind these calculations is
147 cnh 1.2 given later. Indeed many of the illustrative examples shown below can be
148     easily reproduced: simply download the model (the minimum you need is a PC
149 cnh 1.10 running Linux, together with a FORTRAN\ 77 compiler) and follow the examples
150 cnh 1.2 described in detail in the documentation.
151 cnh 1.1
152     \subsection{Global atmosphere: `Held-Suarez' benchmark}
153 afe 1.18 \begin{rawhtml}
154 afe 1.19 <!-- CMIREDIR:atmospheric_example: -->
155 afe 1.18 \end{rawhtml}
156    
157    
158 cnh 1.1
159 cnh 1.7 A novel feature of MITgcm is its ability to simulate, using one basic algorithm,
160     both atmospheric and oceanographic flows at both small and large scales.
161 cnh 1.2
162 cnh 1.7 Figure \ref{fig:eddy_cs} shows an instantaneous plot of the 500$mb$
163 cnh 1.2 temperature field obtained using the atmospheric isomorph of MITgcm run at
164 edhill 1.25 $2.8^{\circ }$ resolution on the cubed sphere. We see cold air over the pole
165 cnh 1.2 (blue) and warm air along an equatorial band (red). Fully developed
166     baroclinic eddies spawned in the northern hemisphere storm track are
167     evident. There are no mountains or land-sea contrast in this calculation,
168     but you can easily put them in. The model is driven by relaxation to a
169     radiative-convective equilibrium profile, following the description set out
170     in Held and Suarez; 1994 designed to test atmospheric hydrodynamical cores -
171     there are no mountains or land-sea contrast.
172    
173 cnh 1.3 %% CNHbegin
174 jmc 1.28 \input{s_overview/text/cubic_eddies_figure}
175 cnh 1.3 %% CNHend
176    
177 cnh 1.2 As described in Adcroft (2001), a `cubed sphere' is used to discretize the
178 cnh 1.10 globe permitting a uniform griding and obviated the need to Fourier filter.
179 cnh 1.2 The `vector-invariant' form of MITgcm supports any orthogonal curvilinear
180     grid, of which the cubed sphere is just one of many choices.
181 cnh 1.1
182 cnh 1.7 Figure \ref{fig:hs_zave_u} shows the 5-year mean, zonally averaged zonal
183     wind from a 20-level configuration of
184 cnh 1.2 the model. It compares favorable with more conventional spatial
185 cnh 1.7 discretization approaches. The two plots show the field calculated using the
186     cube-sphere grid and the flow calculated using a regular, spherical polar
187     latitude-longitude grid. Both grids are supported within the model.
188 cnh 1.1
189 cnh 1.3 %% CNHbegin
190 jmc 1.28 \input{s_overview/text/hs_zave_u_figure}
191 cnh 1.3 %% CNHend
192    
193 cnh 1.2 \subsection{Ocean gyres}
194 afe 1.18 \begin{rawhtml}
195 afe 1.19 <!-- CMIREDIR:oceanic_example: -->
196 afe 1.18 \end{rawhtml}
197     \begin{rawhtml}
198 afe 1.19 <!-- CMIREDIR:ocean_gyres: -->
199 afe 1.18 \end{rawhtml}
200 cnh 1.1
201 cnh 1.2 Baroclinic instability is a ubiquitous process in the ocean, as well as the
202     atmosphere. Ocean eddies play an important role in modifying the
203     hydrographic structure and current systems of the oceans. Coarse resolution
204     models of the oceans cannot resolve the eddy field and yield rather broad,
205     diffusive patterns of ocean currents. But if the resolution of our models is
206     increased until the baroclinic instability process is resolved, numerical
207     solutions of a different and much more realistic kind, can be obtained.
208    
209 edhill 1.25 Figure \ref{fig:ocean-gyres} shows the surface temperature and
210     velocity field obtained from MITgcm run at $\frac{1}{6}^{\circ }$
211     horizontal resolution on a \textit{lat-lon} grid in which the pole has
212     been rotated by $90^{\circ }$ on to the equator (to avoid the
213     converging of meridian in northern latitudes). 21 vertical levels are
214     used in the vertical with a `lopped cell' representation of
215     topography. The development and propagation of anomalously warm and
216     cold eddies can be clearly seen in the Gulf Stream region. The
217     transport of warm water northward by the mean flow of the Gulf Stream
218     is also clearly visible.
219 cnh 1.1
220 cnh 1.3 %% CNHbegin
221 jmc 1.28 \input{s_overview/text/atl6_figure}
222 cnh 1.3 %% CNHend
223    
224    
225 cnh 1.1 \subsection{Global ocean circulation}
226 afe 1.18 \begin{rawhtml}
227 afe 1.19 <!-- CMIREDIR:global_ocean_circulation: -->
228 afe 1.18 \end{rawhtml}
229 cnh 1.1
230 edhill 1.25 Figure \ref{fig:large-scale-circ} (top) shows the pattern of ocean
231     currents at the surface of a $4^{\circ }$ global ocean model run with
232     15 vertical levels. Lopped cells are used to represent topography on a
233     regular \textit{lat-lon} grid extending from $70^{\circ }N$ to
234     $70^{\circ }S$. The model is driven using monthly-mean winds with
235     mixed boundary conditions on temperature and salinity at the surface.
236     The transfer properties of ocean eddies, convection and mixing is
237     parameterized in this model.
238 cnh 1.2
239 cnh 1.7 Figure \ref{fig:large-scale-circ} (bottom) shows the meridional overturning
240     circulation of the global ocean in Sverdrups.
241 cnh 1.2
242 cnh 1.3 %%CNHbegin
243 jmc 1.28 \input{s_overview/text/global_circ_figure}
244 cnh 1.3 %%CNHend
245    
246 cnh 1.2 \subsection{Convection and mixing over topography}
247 afe 1.18 \begin{rawhtml}
248 afe 1.19 <!-- CMIREDIR:mixing_over_topography: -->
249 afe 1.18 \end{rawhtml}
250    
251 cnh 1.2
252     Dense plumes generated by localized cooling on the continental shelf of the
253     ocean may be influenced by rotation when the deformation radius is smaller
254     than the width of the cooling region. Rather than gravity plumes, the
255     mechanism for moving dense fluid down the shelf is then through geostrophic
256 adcroft 1.9 eddies. The simulation shown in the figure \ref{fig:convect-and-topo}
257 cnh 1.7 (blue is cold dense fluid, red is
258 cnh 1.2 warmer, lighter fluid) employs the non-hydrostatic capability of MITgcm to
259     trigger convection by surface cooling. The cold, dense water falls down the
260     slope but is deflected along the slope by rotation. It is found that
261     entrainment in the vertical plane is reduced when rotational control is
262     strong, and replaced by lateral entrainment due to the baroclinic
263     instability of the along-slope current.
264 cnh 1.1
265 cnh 1.3 %%CNHbegin
266 jmc 1.28 \input{s_overview/text/convect_and_topo}
267 cnh 1.3 %%CNHend
268    
269 cnh 1.1 \subsection{Boundary forced internal waves}
270 afe 1.18 \begin{rawhtml}
271 afe 1.19 <!-- CMIREDIR:boundary_forced_internal_waves: -->
272 afe 1.18 \end{rawhtml}
273 cnh 1.1
274 cnh 1.2 The unique ability of MITgcm to treat non-hydrostatic dynamics in the
275     presence of complex geometry makes it an ideal tool to study internal wave
276     dynamics and mixing in oceanic canyons and ridges driven by large amplitude
277     barotropic tidal currents imposed through open boundary conditions.
278    
279 cnh 1.7 Fig. \ref{fig:boundary-forced-wave} shows the influence of cross-slope
280     topographic variations on
281 cnh 1.2 internal wave breaking - the cross-slope velocity is in color, the density
282     contoured. The internal waves are excited by application of open boundary
283 cnh 1.7 conditions on the left. They propagate to the sloping boundary (represented
284 cnh 1.2 using MITgcm's finite volume spatial discretization) where they break under
285     nonhydrostatic dynamics.
286    
287 cnh 1.3 %%CNHbegin
288 jmc 1.28 \input{s_overview/text/boundary_forced_waves}
289 cnh 1.3 %%CNHend
290    
291 cnh 1.2 \subsection{Parameter sensitivity using the adjoint of MITgcm}
292 afe 1.18 \begin{rawhtml}
293 afe 1.19 <!-- CMIREDIR:parameter_sensitivity: -->
294 afe 1.18 \end{rawhtml}
295 cnh 1.2
296     Forward and tangent linear counterparts of MITgcm are supported using an
297     `automatic adjoint compiler'. These can be used in parameter sensitivity and
298     data assimilation studies.
299    
300 edhill 1.25 As one example of application of the MITgcm adjoint, Figure
301     \ref{fig:hf-sensitivity} maps the gradient $\frac{\partial J}{\partial
302     \mathcal{H}}$where $J$ is the magnitude of the overturning
303     stream-function shown in figure \ref{fig:large-scale-circ} at
304     $60^{\circ }N$ and $ \mathcal{H}(\lambda,\varphi)$ is the mean, local
305     air-sea heat flux over a 100 year period. We see that $J$ is sensitive
306     to heat fluxes over the Labrador Sea, one of the important sources of
307     deep water for the thermohaline circulations. This calculation also
308 cnh 1.2 yields sensitivities to all other model parameters.
309    
310 cnh 1.3 %%CNHbegin
311 jmc 1.28 \input{s_overview/text/adj_hf_ocean_figure}
312 cnh 1.3 %%CNHend
313    
314 cnh 1.2 \subsection{Global state estimation of the ocean}
315 afe 1.18 \begin{rawhtml}
316 afe 1.19 <!-- CMIREDIR:global_state_estimation: -->
317 afe 1.18 \end{rawhtml}
318    
319 cnh 1.2
320     An important application of MITgcm is in state estimation of the global
321     ocean circulation. An appropriately defined `cost function', which measures
322     the departure of the model from observations (both remotely sensed and
323 cnh 1.10 in-situ) over an interval of time, is minimized by adjusting `control
324 cnh 1.2 parameters' such as air-sea fluxes, the wind field, the initial conditions
325 cnh 1.15 etc. Figure \ref{fig:assimilated-globes} shows the large scale planetary
326     circulation and a Hopf-Muller plot of Equatorial sea-surface height.
327     Both are obtained from assimilation bringing the model in to
328 cnh 1.7 consistency with altimetric and in-situ observations over the period
329 cnh 1.15 1992-1997.
330 cnh 1.2
331 cnh 1.3 %% CNHbegin
332 jmc 1.28 \input{s_overview/text/assim_figure}
333 cnh 1.3 %% CNHend
334    
335 cnh 1.2 \subsection{Ocean biogeochemical cycles}
336 afe 1.18 \begin{rawhtml}
337 afe 1.19 <!-- CMIREDIR:ocean_biogeo_cycles: -->
338 afe 1.18 \end{rawhtml}
339 cnh 1.2
340 edhill 1.25 MITgcm is being used to study global biogeochemical cycles in the
341     ocean. For example one can study the effects of interannual changes in
342     meteorological forcing and upper ocean circulation on the fluxes of
343     carbon dioxide and oxygen between the ocean and atmosphere. Figure
344     \ref{fig:biogeo} shows the annual air-sea flux of oxygen and its
345     relation to density outcrops in the southern oceans from a single year
346     of a global, interannually varying simulation. The simulation is run
347     at $1^{\circ}\times1^{\circ}$ resolution telescoping to
348     $\frac{1}{3}^{\circ}\times\frac{1}{3}^{\circ}$ in the tropics (not
349     shown).
350 cnh 1.2
351 cnh 1.3 %%CNHbegin
352 jmc 1.28 \input{s_overview/text/biogeo_figure}
353 cnh 1.3 %%CNHend
354 cnh 1.2
355     \subsection{Simulations of laboratory experiments}
356 afe 1.18 \begin{rawhtml}
357 afe 1.19 <!-- CMIREDIR:classroom_exp: -->
358 afe 1.18 \end{rawhtml}
359 cnh 1.2
360 cnh 1.7 Figure \ref{fig:lab-simulation} shows MITgcm being used to simulate a
361 edhill 1.17 laboratory experiment inquiring into the dynamics of the Antarctic Circumpolar Current (ACC). An
362 cnh 1.2 initially homogeneous tank of water ($1m$ in diameter) is driven from its
363     free surface by a rotating heated disk. The combined action of mechanical
364     and thermal forcing creates a lens of fluid which becomes baroclinically
365     unstable. The stratification and depth of penetration of the lens is
366 cnh 1.7 arrested by its instability in a process analogous to that which sets the
367 cnh 1.2 stratification of the ACC.
368 cnh 1.1
369 cnh 1.3 %%CNHbegin
370 jmc 1.28 \input{s_overview/text/lab_figure}
371 cnh 1.3 %%CNHend
372    
373 cnh 1.1 \section{Continuous equations in `r' coordinates}
374 afe 1.18 \begin{rawhtml}
375 afe 1.19 <!-- CMIREDIR:z-p_isomorphism: -->
376 afe 1.18 \end{rawhtml}
377 cnh 1.1
378     To render atmosphere and ocean models from one dynamical core we exploit
379     `isomorphisms' between equation sets that govern the evolution of the
380 cnh 1.7 respective fluids - see figure \ref{fig:isomorphic-equations}.
381     One system of hydrodynamical equations is written down
382 cnh 1.1 and encoded. The model variables have different interpretations depending on
383     whether the atmosphere or ocean is being studied. Thus, for example, the
384     vertical coordinate `$r$' is interpreted as pressure, $p$, if we are
385 edhill 1.17 modeling the atmosphere (right hand side of figure \ref{fig:isomorphic-equations})
386     and height, $z$, if we are modeling the ocean (left hand side of figure
387 cnh 1.7 \ref{fig:isomorphic-equations}).
388 cnh 1.1
389 cnh 1.3 %%CNHbegin
390 jmc 1.28 \input{s_overview/text/zandpcoord_figure.tex}
391 cnh 1.3 %%CNHend
392    
393 cnh 1.1 The state of the fluid at any time is characterized by the distribution of
394     velocity $\vec{\mathbf{v}}$, active tracers $\theta $ and $S$, a
395     `geopotential' $\phi $ and density $\rho =\rho (\theta ,S,p)$ which may
396     depend on $\theta $, $S$, and $p$. The equations that govern the evolution
397     of these fields, obtained by applying the laws of classical mechanics and
398     thermodynamics to a Boussinesq, Navier-Stokes fluid are, written in terms of
399 cnh 1.7 a generic vertical coordinate, $r$, so that the appropriate
400     kinematic boundary conditions can be applied isomorphically
401     see figure \ref{fig:zandp-vert-coord}.
402 cnh 1.1
403 cnh 1.3 %%CNHbegin
404 jmc 1.28 \input{s_overview/text/vertcoord_figure.tex}
405 cnh 1.3 %%CNHend
406    
407 jmc 1.20 \begin{equation}
408 adcroft 1.4 \frac{D\vec{\mathbf{v}_{h}}}{Dt}+\left( 2\vec{\Omega}\times \vec{\mathbf{v}}
409     \right) _{h}+\mathbf{\nabla }_{h}\phi =\mathcal{F}_{\vec{\mathbf{v}_{h}}}
410 cnh 1.8 \text{ horizontal mtm} \label{eq:horizontal_mtm}
411 jmc 1.20 \end{equation}
412 cnh 1.1
413 cnh 1.8 \begin{equation}
414 adcroft 1.4 \frac{D\dot{r}}{Dt}+\widehat{k}\cdot \left( 2\vec{\Omega}\times \vec{\mathbf{
415 cnh 1.1 v}}\right) +\frac{\partial \phi }{\partial r}+b=\mathcal{F}_{\dot{r}}\text{
416 cnh 1.8 vertical mtm} \label{eq:vertical_mtm}
417     \end{equation}
418 cnh 1.1
419     \begin{equation}
420 adcroft 1.4 \mathbf{\nabla }_{h}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \dot{r}}{
421 cnh 1.8 \partial r}=0\text{ continuity} \label{eq:continuity}
422 cnh 1.1 \end{equation}
423    
424 cnh 1.8 \begin{equation}
425     b=b(\theta ,S,r)\text{ equation of state} \label{eq:equation_of_state}
426     \end{equation}
427 cnh 1.1
428 cnh 1.8 \begin{equation}
429 cnh 1.2 \frac{D\theta }{Dt}=\mathcal{Q}_{\theta }\text{ potential temperature}
430 cnh 1.8 \label{eq:potential_temperature}
431     \end{equation}
432 cnh 1.1
433 cnh 1.8 \begin{equation}
434 cnh 1.2 \frac{DS}{Dt}=\mathcal{Q}_{S}\text{ humidity/salinity}
435 adcroft 1.9 \label{eq:humidity_salt}
436 cnh 1.8 \end{equation}
437 cnh 1.1
438     Here:
439    
440     \begin{equation*}
441 cnh 1.2 r\text{ is the vertical coordinate}
442 cnh 1.1 \end{equation*}
443    
444     \begin{equation*}
445     \frac{D}{Dt}=\frac{\partial }{\partial t}+\vec{\mathbf{v}}\cdot \nabla \text{
446 cnh 1.2 is the total derivative}
447 cnh 1.1 \end{equation*}
448    
449     \begin{equation*}
450 adcroft 1.4 \mathbf{\nabla }=\mathbf{\nabla }_{h}+\widehat{k}\frac{\partial }{\partial r}
451 cnh 1.2 \text{ is the `grad' operator}
452 cnh 1.1 \end{equation*}
453 adcroft 1.4 with $\mathbf{\nabla }_{h}$ operating in the horizontal and $\widehat{k}
454 cnh 1.1 \frac{\partial }{\partial r}$ operating in the vertical, where $\widehat{k}$
455     is a unit vector in the vertical
456    
457     \begin{equation*}
458 cnh 1.2 t\text{ is time}
459 cnh 1.1 \end{equation*}
460    
461     \begin{equation*}
462     \vec{\mathbf{v}}=(u,v,\dot{r})=(\vec{\mathbf{v}}_{h},\dot{r})\text{ is the
463 cnh 1.2 velocity}
464 cnh 1.1 \end{equation*}
465    
466     \begin{equation*}
467 cnh 1.2 \phi \text{ is the `pressure'/`geopotential'}
468 cnh 1.1 \end{equation*}
469    
470     \begin{equation*}
471 cnh 1.2 \vec{\Omega}\text{ is the Earth's rotation}
472 cnh 1.1 \end{equation*}
473    
474     \begin{equation*}
475 cnh 1.2 b\text{ is the `buoyancy'}
476 cnh 1.1 \end{equation*}
477    
478     \begin{equation*}
479 cnh 1.2 \theta \text{ is potential temperature}
480 cnh 1.1 \end{equation*}
481    
482     \begin{equation*}
483 cnh 1.2 S\text{ is specific humidity in the atmosphere; salinity in the ocean}
484 cnh 1.1 \end{equation*}
485    
486     \begin{equation*}
487 adcroft 1.4 \mathcal{F}_{\vec{\mathbf{v}}}\text{ are forcing and dissipation of }\vec{
488 cnh 1.1 \mathbf{v}}
489     \end{equation*}
490    
491     \begin{equation*}
492 cnh 1.2 \mathcal{Q}_{\theta }\mathcal{\ }\text{are forcing and dissipation of }\theta
493 cnh 1.1 \end{equation*}
494    
495     \begin{equation*}
496     \mathcal{Q}_{S}\mathcal{\ }\text{are forcing and dissipation of }S
497     \end{equation*}
498    
499     The $\mathcal{F}^{\prime }s$ and $\mathcal{Q}^{\prime }s$ are provided by
500 cnh 1.7 `physics' and forcing packages for atmosphere and ocean. These are described
501     in later chapters.
502 cnh 1.1
503     \subsection{Kinematic Boundary conditions}
504    
505     \subsubsection{vertical}
506    
507 cnh 1.7 at fixed and moving $r$ surfaces we set (see figure \ref{fig:zandp-vert-coord}):
508 cnh 1.1
509     \begin{equation}
510 edhill 1.17 \dot{r}=0 \text{\ at\ } r=R_{fixed}(x,y)\text{ (ocean bottom, top of the atmosphere)}
511 cnh 1.1 \label{eq:fixedbc}
512     \end{equation}
513    
514     \begin{equation}
515 edhill 1.17 \dot{r}=\frac{Dr}{Dt} \text{\ at\ } r=R_{moving}\text{ \
516 cnh 1.10 (ocean surface,bottom of the atmosphere)} \label{eq:movingbc}
517 cnh 1.1 \end{equation}
518    
519     Here
520    
521     \begin{equation*}
522 cnh 1.2 R_{moving}=R_{o}+\eta
523 cnh 1.1 \end{equation*}
524     where $R_{o}(x,y)$ is the `$r-$value' (height or pressure, depending on
525     whether we are in the atmosphere or ocean) of the `moving surface' in the
526     resting fluid and $\eta $ is the departure from $R_{o}(x,y)$ in the presence
527     of motion.
528    
529     \subsubsection{horizontal}
530    
531     \begin{equation}
532     \vec{\mathbf{v}}\cdot \vec{\mathbf{n}}=0 \label{eq:noflow}
533 adcroft 1.4 \end{equation}
534 cnh 1.1 where $\vec{\mathbf{n}}$ is the normal to a solid boundary.
535    
536     \subsection{Atmosphere}
537    
538 cnh 1.7 In the atmosphere, (see figure \ref{fig:zandp-vert-coord}), we interpret:
539 cnh 1.1
540     \begin{equation}
541     r=p\text{ is the pressure} \label{eq:atmos-r}
542     \end{equation}
543    
544     \begin{equation}
545     \dot{r}=\frac{Dp}{Dt}=\omega \text{ is the vertical velocity in }p\text{
546     coordinates} \label{eq:atmos-omega}
547     \end{equation}
548    
549     \begin{equation}
550     \phi =g\,z\text{ is the geopotential height} \label{eq:atmos-phi}
551     \end{equation}
552    
553     \begin{equation}
554     b=\frac{\partial \Pi }{\partial p}\theta \text{ is the buoyancy}
555     \label{eq:atmos-b}
556     \end{equation}
557    
558     \begin{equation}
559     \theta =T(\frac{p_{c}}{p})^{\kappa }\text{ is potential temperature}
560     \label{eq:atmos-theta}
561     \end{equation}
562    
563     \begin{equation}
564     S=q,\text{ is the specific humidity} \label{eq:atmos-s}
565     \end{equation}
566     where
567    
568     \begin{equation*}
569     T\text{ is absolute temperature}
570 adcroft 1.4 \end{equation*}
571 cnh 1.1 \begin{equation*}
572     p\text{ is the pressure}
573 adcroft 1.4 \end{equation*}
574 cnh 1.1 \begin{eqnarray*}
575     &&z\text{ is the height of the pressure surface} \\
576     &&g\text{ is the acceleration due to gravity}
577     \end{eqnarray*}
578    
579     In the above the ideal gas law, $p=\rho RT$, has been expressed in terms of
580     the Exner function $\Pi (p)$ given by (see Appendix Atmosphere)
581     \begin{equation}
582     \Pi (p)=c_{p}(\frac{p}{p_{c}})^{\kappa } \label{eq:exner}
583 adcroft 1.4 \end{equation}
584 cnh 1.1 where $p_{c}$ is a reference pressure and $\kappa =R/c_{p}$ with $R$ the gas
585     constant and $c_{p}$ the specific heat of air at constant pressure.
586    
587     At the top of the atmosphere (which is `fixed' in our $r$ coordinate):
588    
589     \begin{equation*}
590 cnh 1.2 R_{fixed}=p_{top}=0
591 cnh 1.1 \end{equation*}
592     In a resting atmosphere the elevation of the mountains at the bottom is
593     given by
594     \begin{equation*}
595 cnh 1.2 R_{moving}=R_{o}(x,y)=p_{o}(x,y)
596 cnh 1.1 \end{equation*}
597     i.e. the (hydrostatic) pressure at the top of the mountains in a resting
598     atmosphere.
599    
600     The boundary conditions at top and bottom are given by:
601    
602     \begin{eqnarray}
603     &&\omega =0~\text{at }r=R_{fixed} \text{ (top of the atmosphere)}
604     \label{eq:fixed-bc-atmos} \\
605     \omega &=&\frac{Dp_{s}}{Dt}\text{; at }r=R_{moving}\text{ (bottom of the
606     atmosphere)} \label{eq:moving-bc-atmos}
607     \end{eqnarray}
608    
609 edhill 1.21 Then the (hydrostatic form of) equations
610     (\ref{eq:horizontal_mtm}-\ref{eq:humidity_salt}) yields a consistent
611     set of atmospheric equations which, for convenience, are written out
612     in $p$ coordinates in Appendix Atmosphere - see
613     eqs(\ref{eq:atmos-prime}).
614 cnh 1.1
615     \subsection{Ocean}
616    
617     In the ocean we interpret:
618     \begin{eqnarray}
619     r &=&z\text{ is the height} \label{eq:ocean-z} \\
620     \dot{r} &=&\frac{Dz}{Dt}=w\text{ is the vertical velocity}
621     \label{eq:ocean-w} \\
622     \phi &=&\frac{p}{\rho _{c}}\text{ is the pressure} \label{eq:ocean-p} \\
623     b(\theta ,S,r) &=&\frac{g}{\rho _{c}}\left( \rho (\theta ,S,r)-\rho
624     _{c}\right) \text{ is the buoyancy} \label{eq:ocean-b}
625     \end{eqnarray}
626     where $\rho _{c}$ is a fixed reference density of water and $g$ is the
627     acceleration due to gravity.\noindent
628    
629     In the above
630    
631     At the bottom of the ocean: $R_{fixed}(x,y)=-H(x,y)$.
632    
633     The surface of the ocean is given by: $R_{moving}=\eta $
634    
635 adcroft 1.4 The position of the resting free surface of the ocean is given by $
636 cnh 1.1 R_{o}=Z_{o}=0$.
637    
638     Boundary conditions are:
639    
640     \begin{eqnarray}
641     w &=&0~\text{at }r=R_{fixed}\text{ (ocean bottom)} \label{eq:fixed-bc-ocean}
642     \\
643 adcroft 1.4 w &=&\frac{D\eta }{Dt}\text{ at }r=R_{moving}=\eta \text{ (ocean surface)
644 cnh 1.1 \label{eq:moving-bc-ocean}}
645     \end{eqnarray}
646     where $\eta $ is the elevation of the free surface.
647    
648 adcroft 1.9 Then equations (\ref{eq:horizontal_mtm}-\ref{eq:humidity_salt}) yield a consistent set
649 cnh 1.8 of oceanic equations
650 cnh 1.1 which, for convenience, are written out in $z$ coordinates in Appendix Ocean
651     - see eqs(\ref{eq:ocean-mom}) to (\ref{eq:ocean-salt}).
652    
653     \subsection{Hydrostatic, Quasi-hydrostatic, Quasi-nonhydrostatic and
654     Non-hydrostatic forms}
655 jmc 1.29 \label{sec:all_hydrostatic_forms}
656 afe 1.18 \begin{rawhtml}
657 afe 1.19 <!-- CMIREDIR:non_hydrostatic: -->
658 afe 1.18 \end{rawhtml}
659    
660 cnh 1.1
661     Let us separate $\phi $ in to surface, hydrostatic and non-hydrostatic terms:
662    
663     \begin{equation}
664     \phi (x,y,r)=\phi _{s}(x,y)+\phi _{hyd}(x,y,r)+\phi _{nh}(x,y,r)
665     \label{eq:phi-split}
666 adcroft 1.4 \end{equation}
667 jmc 1.20 %and write eq(\ref{eq:incompressible}) in the form:
668     % ^- this eq is missing (jmc) ; replaced with:
669     and write eq( \ref{eq:horizontal_mtm}) in the form:
670 cnh 1.1
671     \begin{equation}
672     \frac{\partial \vec{\mathbf{v}_{h}}}{\partial t}+\mathbf{\nabla }_{h}\phi
673     _{s}+\mathbf{\nabla }_{h}\phi _{hyd}+\epsilon _{nh}\mathbf{\nabla }_{h}\phi
674     _{nh}=\vec{\mathbf{G}}_{\vec{v}_{h}} \label{eq:mom-h}
675     \end{equation}
676    
677     \begin{equation}
678     \frac{\partial \phi _{hyd}}{\partial r}=-b \label{eq:hydrostatic}
679     \end{equation}
680    
681     \begin{equation}
682 adcroft 1.4 \epsilon _{nh}\frac{\partial \dot{r}}{\partial t}+\frac{\partial \phi _{nh}}{
683 cnh 1.1 \partial r}=G_{\dot{r}} \label{eq:mom-w}
684     \end{equation}
685     Here $\epsilon _{nh}$ is a non-hydrostatic parameter.
686    
687 adcroft 1.4 The $\left( \vec{\mathbf{G}}_{\vec{v}},G_{\dot{r}}\right) $ in eq(\ref
688 cnh 1.1 {eq:mom-h}) and (\ref{eq:mom-w}) represent advective, metric and Coriolis
689 adcroft 1.4 terms in the momentum equations. In spherical coordinates they take the form
690     \footnote{
691 cnh 1.1 In the hydrostatic primitive equations (\textbf{HPE}) all underlined terms
692 adcroft 1.4 in (\ref{eq:gu-speherical}), (\ref{eq:gv-spherical}) and (\ref
693 cnh 1.1 {eq:gw-spherical}) are omitted; the singly-underlined terms are included in
694 adcroft 1.4 the quasi-hydrostatic model (\textbf{QH}). The fully non-hydrostatic model (
695 cnh 1.1 \textbf{NH}) includes all terms.} - see Marshall et al 1997a for a full
696     discussion:
697    
698     \begin{equation}
699     \left.
700     \begin{tabular}{l}
701     $G_{u}=-\vec{\mathbf{v}}.\nabla u$ \\
702 cnh 1.6 $-\left\{ \underline{\frac{u\dot{r}}{{r}}}-\frac{uv\tan \varphi}{{r}}\right\} $
703 cnh 1.1 \\
704 cnh 1.6 $-\left\{ -2\Omega v\sin \varphi+\underline{2\Omega \dot{r}\cos \varphi}\right\} $
705 cnh 1.1 \\
706 adcroft 1.4 $+\mathcal{F}_{u}$
707     \end{tabular}
708 cnh 1.1 \ \right\} \left\{
709     \begin{tabular}{l}
710     \textit{advection} \\
711     \textit{metric} \\
712     \textit{Coriolis} \\
713 adcroft 1.4 \textit{\ Forcing/Dissipation}
714     \end{tabular}
715 cnh 1.2 \ \right. \qquad \label{eq:gu-speherical}
716 cnh 1.1 \end{equation}
717    
718     \begin{equation}
719     \left.
720     \begin{tabular}{l}
721     $G_{v}=-\vec{\mathbf{v}}.\nabla v$ \\
722 cnh 1.6 $-\left\{ \underline{\frac{v\dot{r}}{{r}}}-\frac{u^{2}\tan \varphi}{{r}}\right\}
723 cnh 1.1 $ \\
724 cnh 1.6 $-\left\{ -2\Omega u\sin \varphi \right\} $ \\
725 adcroft 1.4 $+\mathcal{F}_{v}$
726     \end{tabular}
727 cnh 1.1 \ \right\} \left\{
728     \begin{tabular}{l}
729     \textit{advection} \\
730     \textit{metric} \\
731     \textit{Coriolis} \\
732 adcroft 1.4 \textit{\ Forcing/Dissipation}
733     \end{tabular}
734 cnh 1.2 \ \right. \qquad \label{eq:gv-spherical}
735 adcroft 1.4 \end{equation}
736 cnh 1.2 \qquad \qquad \qquad \qquad \qquad
737 cnh 1.1
738     \begin{equation}
739     \left.
740     \begin{tabular}{l}
741     $G_{\dot{r}}=-\underline{\underline{\vec{\mathbf{v}}.\nabla \dot{r}}}$ \\
742     $+\left\{ \underline{\frac{u^{_{^{2}}}+v^{2}}{{r}}}\right\} $ \\
743 cnh 1.6 ${+}\underline{{2\Omega u\cos \varphi}}$ \\
744 adcroft 1.4 $\underline{\underline{\mathcal{F}_{\dot{r}}}}$
745     \end{tabular}
746 cnh 1.1 \ \right\} \left\{
747     \begin{tabular}{l}
748     \textit{advection} \\
749     \textit{metric} \\
750     \textit{Coriolis} \\
751 adcroft 1.4 \textit{\ Forcing/Dissipation}
752     \end{tabular}
753 cnh 1.2 \ \right. \label{eq:gw-spherical}
754 adcroft 1.4 \end{equation}
755 cnh 1.2 \qquad \qquad \qquad \qquad \qquad
756 cnh 1.1
757 cnh 1.6 In the above `${r}$' is the distance from the center of the earth and `$\varphi$
758 cnh 1.1 ' is latitude.
759    
760     Grad and div operators in spherical coordinates are defined in appendix
761 adcroft 1.4 OPERATORS.
762 cnh 1.1
763 cnh 1.3 %%CNHbegin
764 jmc 1.28 \input{s_overview/text/sphere_coord_figure.tex}
765 cnh 1.3 %%CNHend
766    
767 cnh 1.1 \subsubsection{Shallow atmosphere approximation}
768    
769 edhill 1.24 Most models are based on the `hydrostatic primitive equations' (HPE's)
770     in which the vertical momentum equation is reduced to a statement of
771     hydrostatic balance and the `traditional approximation' is made in
772     which the Coriolis force is treated approximately and the shallow
773     atmosphere approximation is made. MITgcm need not make the
774     `traditional approximation'. To be able to support consistent
775     non-hydrostatic forms the shallow atmosphere approximation can be
776     relaxed - when dividing through by $ r $ in, for example,
777     (\ref{eq:gu-speherical}), we do not replace $r$ by $a$, the radius of
778     the earth.
779 cnh 1.1
780     \subsubsection{Hydrostatic and quasi-hydrostatic forms}
781 cnh 1.7 \label{sec:hydrostatic_and_quasi-hydrostatic_forms}
782 cnh 1.1
783     These are discussed at length in Marshall et al (1997a).
784    
785     In the `hydrostatic primitive equations' (\textbf{HPE)} all the underlined
786     terms in Eqs. (\ref{eq:gu-speherical} $\rightarrow $\ \ref{eq:gw-spherical})
787     are neglected and `${r}$' is replaced by `$a$', the mean radius of the
788     earth. Once the pressure is found at one level - e.g. by inverting a 2-d
789     Elliptic equation for $\phi _{s}$ at $r=R_{moving}$ - the pressure can be
790 adcroft 1.4 computed at all other levels by integration of the hydrostatic relation, eq(
791 cnh 1.1 \ref{eq:hydrostatic}).
792    
793     In the `quasi-hydrostatic' equations (\textbf{QH)} strict balance between
794     gravity and vertical pressure gradients is not imposed. The $2\Omega u\cos
795 cnh 1.6 \varphi $ Coriolis term are not neglected and are balanced by a non-hydrostatic
796 adcroft 1.4 contribution to the pressure field: only the terms underlined twice in Eqs. (
797 cnh 1.1 \ref{eq:gu-speherical}$\rightarrow $\ \ref{eq:gw-spherical}) are set to zero
798     and, simultaneously, the shallow atmosphere approximation is relaxed. In
799     \textbf{QH}\ \textit{all} the metric terms are retained and the full
800     variation of the radial position of a particle monitored. The \textbf{QH}\
801     vertical momentum equation (\ref{eq:mom-w}) becomes:
802    
803     \begin{equation*}
804 cnh 1.6 \frac{\partial \phi _{nh}}{\partial r}=2\Omega u\cos \varphi
805 cnh 1.1 \end{equation*}
806     making a small correction to the hydrostatic pressure.
807    
808     \textbf{QH} has good energetic credentials - they are the same as for
809     \textbf{HPE}. Importantly, however, it has the same angular momentum
810     principle as the full non-hydrostatic model (\textbf{NH)} - see Marshall
811     et.al., 1997a. As in \textbf{HPE }only a 2-d elliptic problem need be solved.
812    
813     \subsubsection{Non-hydrostatic and quasi-nonhydrostatic forms}
814    
815 edhill 1.24 MITgcm presently supports a full non-hydrostatic ocean isomorph, but
816 cnh 1.1 only a quasi-non-hydrostatic atmospheric isomorph.
817    
818     \paragraph{Non-hydrostatic Ocean}
819    
820 adcroft 1.4 In the non-hydrostatic ocean model all terms in equations Eqs.(\ref
821 cnh 1.1 {eq:gu-speherical} $\rightarrow $\ \ref{eq:gw-spherical}) are retained. A
822     three dimensional elliptic equation must be solved subject to Neumann
823     boundary conditions (see below). It is important to note that use of the
824     full \textbf{NH} does not admit any new `fast' waves in to the system - the
825 cnh 1.8 incompressible condition eq(\ref{eq:continuity}) has already filtered out
826 cnh 1.1 acoustic modes. It does, however, ensure that the gravity waves are treated
827     accurately with an exact dispersion relation. The \textbf{NH} set has a
828     complete angular momentum principle and consistent energetics - see White
829     and Bromley, 1995; Marshall et.al.\ 1997a.
830    
831     \paragraph{Quasi-nonhydrostatic Atmosphere}
832    
833 adcroft 1.4 In the non-hydrostatic version of our atmospheric model we approximate $\dot{
834 cnh 1.1 r}$ in the vertical momentum eqs(\ref{eq:mom-w}) and (\ref{eq:gv-spherical})
835     (but only here) by:
836    
837     \begin{equation}
838     \dot{r}=\frac{Dp}{Dt}=\frac{1}{g}\frac{D\phi }{Dt} \label{eq:quasi-nh-w}
839 adcroft 1.4 \end{equation}
840 cnh 1.1 where $p_{hy}$ is the hydrostatic pressure.
841    
842     \subsubsection{Summary of equation sets supported by model}
843    
844     \paragraph{Atmosphere}
845    
846     Hydrostatic, and quasi-hydrostatic and quasi non-hydrostatic forms of the
847     compressible non-Boussinesq equations in $p-$coordinates are supported.
848    
849     \subparagraph{Hydrostatic and quasi-hydrostatic}
850    
851     The hydrostatic set is written out in $p-$coordinates in appendix Atmosphere
852     - see eq(\ref{eq:atmos-prime}).
853    
854     \subparagraph{Quasi-nonhydrostatic}
855    
856     A quasi-nonhydrostatic form is also supported.
857    
858     \paragraph{Ocean}
859    
860     \subparagraph{Hydrostatic and quasi-hydrostatic}
861    
862     Hydrostatic, and quasi-hydrostatic forms of the incompressible Boussinesq
863     equations in $z-$coordinates are supported.
864    
865     \subparagraph{Non-hydrostatic}
866    
867 adcroft 1.4 Non-hydrostatic forms of the incompressible Boussinesq equations in $z-$
868     coordinates are supported - see eqs(\ref{eq:ocean-mom}) to (\ref
869 cnh 1.1 {eq:ocean-salt}).
870    
871     \subsection{Solution strategy}
872    
873 adcroft 1.4 The method of solution employed in the \textbf{HPE}, \textbf{QH} and \textbf{
874 cnh 1.8 NH} models is summarized in Figure \ref{fig:solution-strategy}.
875     Under all dynamics, a 2-d elliptic equation is
876 cnh 1.1 first solved to find the surface pressure and the hydrostatic pressure at
877     any level computed from the weight of fluid above. Under \textbf{HPE} and
878     \textbf{QH} dynamics, the horizontal momentum equations are then stepped
879     forward and $\dot{r}$ found from continuity. Under \textbf{NH} dynamics a
880     3-d elliptic equation must be solved for the non-hydrostatic pressure before
881     stepping forward the horizontal momentum equations; $\dot{r}$ is found by
882     stepping forward the vertical momentum equation.
883    
884 cnh 1.3 %%CNHbegin
885 jmc 1.28 \input{s_overview/text/solution_strategy_figure.tex}
886 cnh 1.3 %%CNHend
887    
888 cnh 1.1 There is no penalty in implementing \textbf{QH} over \textbf{HPE} except, of
889 cnh 1.6 course, some complication that goes with the inclusion of $\cos \varphi \ $
890 cnh 1.1 Coriolis terms and the relaxation of the shallow atmosphere approximation.
891     But this leads to negligible increase in computation. In \textbf{NH}, in
892     contrast, one additional elliptic equation - a three-dimensional one - must
893     be inverted for $p_{nh}$. However the `overhead' of the \textbf{NH} model is
894     essentially negligible in the hydrostatic limit (see detailed discussion in
895     Marshall et al, 1997) resulting in a non-hydrostatic algorithm that, in the
896     hydrostatic limit, is as computationally economic as the \textbf{HPEs}.
897    
898     \subsection{Finding the pressure field}
899 cnh 1.7 \label{sec:finding_the_pressure_field}
900 cnh 1.1
901     Unlike the prognostic variables $u$, $v$, $w$, $\theta $ and $S$, the
902     pressure field must be obtained diagnostically. We proceed, as before, by
903     dividing the total (pressure/geo) potential in to three parts, a surface
904     part, $\phi _{s}(x,y)$, a hydrostatic part $\phi _{hyd}(x,y,r)$ and a
905     non-hydrostatic part $\phi _{nh}(x,y,r)$, as in (\ref{eq:phi-split}), and
906     writing the momentum equation as in (\ref{eq:mom-h}).
907    
908     \subsubsection{Hydrostatic pressure}
909    
910     Hydrostatic pressure is obtained by integrating (\ref{eq:hydrostatic})
911     vertically from $r=R_{o}$ where $\phi _{hyd}(r=R_{o})=0$, to yield:
912    
913     \begin{equation*}
914 adcroft 1.4 \int_{r}^{R_{o}}\frac{\partial \phi _{hyd}}{\partial r}dr=\left[ \phi _{hyd}
915 cnh 1.2 \right] _{r}^{R_{o}}=\int_{r}^{R_{o}}-bdr
916 cnh 1.1 \end{equation*}
917     and so
918    
919     \begin{equation}
920     \phi _{hyd}(x,y,r)=\int_{r}^{R_{o}}bdr \label{eq:hydro-phi}
921     \end{equation}
922    
923     The model can be easily modified to accommodate a loading term (e.g
924     atmospheric pressure pushing down on the ocean's surface) by setting:
925    
926     \begin{equation}
927     \phi _{hyd}(r=R_{o})=loading \label{eq:loading}
928     \end{equation}
929    
930     \subsubsection{Surface pressure}
931    
932 cnh 1.8 The surface pressure equation can be obtained by integrating continuity,
933     (\ref{eq:continuity}), vertically from $r=R_{fixed}$ to $r=R_{moving}$
934 cnh 1.1
935     \begin{equation*}
936 adcroft 1.4 \int_{R_{fixed}}^{R_{moving}}\left( \mathbf{\nabla }_{h}\cdot \vec{\mathbf{v}
937 cnh 1.2 }_{h}+\partial _{r}\dot{r}\right) dr=0
938 cnh 1.1 \end{equation*}
939    
940     Thus:
941    
942     \begin{equation*}
943     \frac{\partial \eta }{\partial t}+\vec{\mathbf{v}}.\nabla \eta
944 adcroft 1.4 +\int_{R_{fixed}}^{R_{moving}}\mathbf{\nabla }_{h}\cdot \vec{\mathbf{v}}
945 cnh 1.2 _{h}dr=0
946 cnh 1.1 \end{equation*}
947 adcroft 1.4 where $\eta =R_{moving}-R_{o}$ is the free-surface $r$-anomaly in units of $
948 cnh 1.1 r $. The above can be rearranged to yield, using Leibnitz's theorem:
949    
950     \begin{equation}
951     \frac{\partial \eta }{\partial t}+\mathbf{\nabla }_{h}\cdot
952     \int_{R_{fixed}}^{R_{moving}}\vec{\mathbf{v}}_{h}dr=\text{source}
953     \label{eq:free-surface}
954 adcroft 1.4 \end{equation}
955 cnh 1.1 where we have incorporated a source term.
956    
957     Whether $\phi $ is pressure (ocean model, $p/\rho _{c}$) or geopotential
958 cnh 1.8 (atmospheric model), in (\ref{eq:mom-h}), the horizontal gradient term can
959 cnh 1.1 be written
960     \begin{equation}
961 cnh 1.2 \mathbf{\nabla }_{h}\phi _{s}=\mathbf{\nabla }_{h}\left( b_{s}\eta \right)
962 cnh 1.1 \label{eq:phi-surf}
963 adcroft 1.4 \end{equation}
964 cnh 1.1 where $b_{s}$ is the buoyancy at the surface.
965    
966 cnh 1.8 In the hydrostatic limit ($\epsilon _{nh}=0$), equations (\ref{eq:mom-h}), (\ref
967 cnh 1.1 {eq:free-surface}) and (\ref{eq:phi-surf}) can be solved by inverting a 2-d
968     elliptic equation for $\phi _{s}$ as described in Chapter 2. Both `free
969     surface' and `rigid lid' approaches are available.
970    
971     \subsubsection{Non-hydrostatic pressure}
972    
973 cnh 1.8 Taking the horizontal divergence of (\ref{eq:mom-h}) and adding
974     $\frac{\partial }{\partial r}$ of (\ref{eq:mom-w}), invoking the continuity equation
975     (\ref{eq:continuity}), we deduce that:
976 cnh 1.1
977     \begin{equation}
978 adcroft 1.4 \nabla _{3}^{2}\phi _{nh}=\nabla .\vec{\mathbf{G}}_{\vec{v}}-\left( \mathbf{
979     \nabla }_{h}^{2}\phi _{s}+\mathbf{\nabla }^{2}\phi _{hyd}\right) =\nabla .
980 cnh 1.1 \vec{\mathbf{F}} \label{eq:3d-invert}
981     \end{equation}
982    
983     For a given rhs this 3-d elliptic equation must be inverted for $\phi _{nh}$
984     subject to appropriate choice of boundary conditions. This method is usually
985     called \textit{The Pressure Method} [Harlow and Welch, 1965; Williams, 1969;
986     Potter, 1976]. In the hydrostatic primitive equations case (\textbf{HPE}),
987     the 3-d problem does not need to be solved.
988    
989     \paragraph{Boundary Conditions}
990    
991     We apply the condition of no normal flow through all solid boundaries - the
992     coasts (in the ocean) and the bottom:
993    
994     \begin{equation}
995     \vec{\mathbf{v}}.\widehat{n}=0 \label{nonormalflow}
996     \end{equation}
997     where $\widehat{n}$ is a vector of unit length normal to the boundary. The
998     kinematic condition (\ref{nonormalflow}) is also applied to the vertical
999 adcroft 1.4 velocity at $r=R_{moving}$. No-slip $\left( v_{T}=0\right) \ $or slip $
1000 cnh 1.1 \left( \partial v_{T}/\partial n=0\right) \ $conditions are employed on the
1001     tangential component of velocity, $v_{T}$, at all solid boundaries,
1002     depending on the form chosen for the dissipative terms in the momentum
1003     equations - see below.
1004    
1005 cnh 1.8 Eq.(\ref{nonormalflow}) implies, making use of (\ref{eq:mom-h}), that:
1006 cnh 1.1
1007     \begin{equation}
1008     \widehat{n}.\nabla \phi _{nh}=\widehat{n}.\vec{\mathbf{F}}
1009     \label{eq:inhom-neumann-nh}
1010     \end{equation}
1011     where
1012    
1013     \begin{equation*}
1014     \vec{\mathbf{F}}=\vec{\mathbf{G}}_{\vec{v}}-\left( \mathbf{\nabla }_{h}\phi
1015     _{s}+\mathbf{\nabla }\phi _{hyd}\right)
1016 adcroft 1.4 \end{equation*}
1017 cnh 1.1 presenting inhomogeneous Neumann boundary conditions to the Elliptic problem
1018     (\ref{eq:3d-invert}). As shown, for example, by Williams (1969), one can
1019     exploit classical 3D potential theory and, by introducing an appropriately
1020 cnh 1.2 chosen $\delta $-function sheet of `source-charge', replace the
1021     inhomogeneous boundary condition on pressure by a homogeneous one. The
1022 adcroft 1.4 source term $rhs$ in (\ref{eq:3d-invert}) is the divergence of the vector $
1023     \vec{\mathbf{F}}.$ By simultaneously setting $
1024 cnh 1.1 \begin{array}{l}
1025 adcroft 1.4 \widehat{n}.\vec{\mathbf{F}}
1026     \end{array}
1027 cnh 1.1 =0$\ and $\widehat{n}.\nabla \phi _{nh}=0\ $on the boundary the following
1028 cnh 1.2 self-consistent but simpler homogenized Elliptic problem is obtained:
1029 cnh 1.1
1030     \begin{equation*}
1031 cnh 1.2 \nabla ^{2}\phi _{nh}=\nabla .\widetilde{\vec{\mathbf{F}}}\qquad
1032 adcroft 1.4 \end{equation*}
1033 cnh 1.1 where $\widetilde{\vec{\mathbf{F}}}$ is a modified $\vec{\mathbf{F}}$ such
1034 adcroft 1.4 that $\widetilde{\vec{\mathbf{F}}}.\widehat{n}=0$. As is implied by (\ref
1035 cnh 1.1 {eq:inhom-neumann-nh}) the modified boundary condition becomes:
1036    
1037     \begin{equation}
1038     \widehat{n}.\nabla \phi _{nh}=0 \label{eq:hom-neumann-nh}
1039     \end{equation}
1040    
1041     If the flow is `close' to hydrostatic balance then the 3-d inversion
1042     converges rapidly because $\phi _{nh}\ $is then only a small correction to
1043     the hydrostatic pressure field (see the discussion in Marshall et al, a,b).
1044    
1045 cnh 1.8 The solution $\phi _{nh}\ $to (\ref{eq:3d-invert}) and (\ref{eq:inhom-neumann-nh})
1046 cnh 1.1 does not vanish at $r=R_{moving}$, and so refines the pressure there.
1047    
1048     \subsection{Forcing/dissipation}
1049    
1050     \subsubsection{Forcing}
1051    
1052     The forcing terms $\mathcal{F}$ on the rhs of the equations are provided by
1053 cnh 1.8 `physics packages' and forcing packages. These are described later on.
1054 cnh 1.1
1055     \subsubsection{Dissipation}
1056    
1057     \paragraph{Momentum}
1058    
1059     Many forms of momentum dissipation are available in the model. Laplacian and
1060     biharmonic frictions are commonly used:
1061    
1062     \begin{equation}
1063 adcroft 1.4 D_{V}=A_{h}\nabla _{h}^{2}v+A_{v}\frac{\partial ^{2}v}{\partial z^{2}}
1064 cnh 1.1 +A_{4}\nabla _{h}^{4}v \label{eq:dissipation}
1065     \end{equation}
1066     where $A_{h}$ and $A_{v}\ $are (constant) horizontal and vertical viscosity
1067     coefficients and $A_{4}\ $is the horizontal coefficient for biharmonic
1068     friction. These coefficients are the same for all velocity components.
1069    
1070     \paragraph{Tracers}
1071    
1072     The mixing terms for the temperature and salinity equations have a similar
1073     form to that of momentum except that the diffusion tensor can be
1074 edhill 1.26 non-diagonal and have varying coefficients.
1075 cnh 1.1 \begin{equation}
1076     D_{T,S}=\nabla .[\underline{\underline{K}}\nabla (T,S)]+K_{4}\nabla
1077     _{h}^{4}(T,S) \label{eq:diffusion}
1078     \end{equation}
1079 adcroft 1.4 where $\underline{\underline{K}}\ $is the diffusion tensor and the $K_{4}\ $
1080 cnh 1.1 horizontal coefficient for biharmonic diffusion. In the simplest case where
1081     the subgrid-scale fluxes of heat and salt are parameterized with constant
1082     horizontal and vertical diffusion coefficients, $\underline{\underline{K}}$,
1083     reduces to a diagonal matrix with constant coefficients:
1084    
1085     \begin{equation}
1086     \qquad \qquad \qquad \qquad K=\left(
1087     \begin{array}{ccc}
1088     K_{h} & 0 & 0 \\
1089     0 & K_{h} & 0 \\
1090 adcroft 1.4 0 & 0 & K_{v}
1091 cnh 1.1 \end{array}
1092     \right) \qquad \qquad \qquad \label{eq:diagonal-diffusion-tensor}
1093     \end{equation}
1094     where $K_{h}\ $and $K_{v}\ $are the horizontal and vertical diffusion
1095     coefficients. These coefficients are the same for all tracers (temperature,
1096     salinity ... ).
1097    
1098     \subsection{Vector invariant form}
1099    
1100 edhill 1.21 For some purposes it is advantageous to write momentum advection in
1101     eq(\ref {eq:horizontal_mtm}) and (\ref{eq:vertical_mtm}) in the
1102     (so-called) `vector invariant' form:
1103 cnh 1.1
1104     \begin{equation}
1105 adcroft 1.4 \frac{D\vec{\mathbf{v}}}{Dt}=\frac{\partial \vec{\mathbf{v}}}{\partial t}
1106     +\left( \nabla \times \vec{\mathbf{v}}\right) \times \vec{\mathbf{v}}+\nabla
1107 cnh 1.2 \left[ \frac{1}{2}(\vec{\mathbf{v}}\cdot \vec{\mathbf{v}})\right]
1108 cnh 1.1 \label{eq:vi-identity}
1109 adcroft 1.4 \end{equation}
1110 cnh 1.1 This permits alternative numerical treatments of the non-linear terms based
1111     on their representation as a vorticity flux. Because gradients of coordinate
1112     vectors no longer appear on the rhs of (\ref{eq:vi-identity}), explicit
1113 adcroft 1.4 representation of the metric terms in (\ref{eq:gu-speherical}), (\ref
1114 cnh 1.1 {eq:gv-spherical}) and (\ref{eq:gw-spherical}), can be avoided: information
1115     about the geometry is contained in the areas and lengths of the volumes used
1116     to discretize the model.
1117    
1118     \subsection{Adjoint}
1119    
1120 cnh 1.8 Tangent linear and adjoint counterparts of the forward model are described
1121 cnh 1.2 in Chapter 5.
1122 cnh 1.1
1123     \section{Appendix ATMOSPHERE}
1124    
1125     \subsection{Hydrostatic Primitive Equations for the Atmosphere in pressure
1126     coordinates}
1127    
1128     \label{sect-hpe-p}
1129    
1130     The hydrostatic primitive equations (HPEs) in p-coordinates are:
1131     \begin{eqnarray}
1132 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1133 cnh 1.2 _{h}+\mathbf{\nabla }_{p}\phi &=&\vec{\mathbf{\mathcal{F}}}
1134 cnh 1.1 \label{eq:atmos-mom} \\
1135 cnh 1.2 \frac{\partial \phi }{\partial p}+\alpha &=&0 \label{eq-p-hydro-start} \\
1136 adcroft 1.4 \mathbf{\nabla }_{p}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \omega }{
1137 cnh 1.1 \partial p} &=&0 \label{eq:atmos-cont} \\
1138 cnh 1.2 p\alpha &=&RT \label{eq:atmos-eos} \\
1139 cnh 1.1 c_{v}\frac{DT}{Dt}+p\frac{D\alpha }{Dt} &=&\mathcal{Q} \label{eq:atmos-heat}
1140 adcroft 1.4 \end{eqnarray}
1141 cnh 1.1 where $\vec{\mathbf{v}}_{h}=(u,v,0)$ is the `horizontal' (on pressure
1142 jmc 1.30 surfaces) component of velocity, $\frac{D}{Dt}=\frac{\partial}{\partial t}
1143     +\vec{\mathbf{v}}_{h}\cdot \mathbf{\nabla }_{p}+\omega \frac{\partial }{\partial p}$
1144     is the total derivative, $f=2\Omega \sin \varphi$ is the Coriolis parameter,
1145     $\phi =gz$ is the geopotential, $\alpha =1/\rho $ is the specific volume,
1146     $\omega =\frac{Dp }{Dt}$ is the vertical velocity in the $p-$coordinate.
1147     Equation(\ref {eq:atmos-heat}) is the first law of thermodynamics where internal
1148     energy $e=c_{v}T$, $T$ is temperature, $Q$ is the rate of heating per unit mass
1149     and $p\frac{D\alpha }{Dt}$ is the work done by the fluid in compressing.
1150 cnh 1.1
1151     It is convenient to cast the heat equation in terms of potential temperature
1152     $\theta $ so that it looks more like a generic conservation law.
1153     Differentiating (\ref{eq:atmos-eos}) we get:
1154     \begin{equation*}
1155     p\frac{D\alpha }{Dt}+\alpha \frac{Dp}{Dt}=R\frac{DT}{Dt}
1156 adcroft 1.4 \end{equation*}
1157     which, when added to the heat equation (\ref{eq:atmos-heat}) and using $
1158 cnh 1.1 c_{p}=c_{v}+R$, gives:
1159     \begin{equation}
1160     c_{p}\frac{DT}{Dt}-\alpha \frac{Dp}{Dt}=\mathcal{Q}
1161     \label{eq-p-heat-interim}
1162 adcroft 1.4 \end{equation}
1163 cnh 1.1 Potential temperature is defined:
1164     \begin{equation}
1165     \theta =T(\frac{p_{c}}{p})^{\kappa } \label{eq:potential-temp}
1166 adcroft 1.4 \end{equation}
1167 cnh 1.1 where $p_{c}$ is a reference pressure and $\kappa =R/c_{p}$. For convenience
1168     we will make use of the Exner function $\Pi (p)$ which defined by:
1169     \begin{equation}
1170     \Pi (p)=c_{p}(\frac{p}{p_{c}})^{\kappa } \label{Exner}
1171 adcroft 1.4 \end{equation}
1172 cnh 1.1 The following relations will be useful and are easily expressed in terms of
1173     the Exner function:
1174     \begin{equation*}
1175     c_{p}T=\Pi \theta \;\;;\;\;\frac{\partial \Pi }{\partial p}=\frac{\kappa \Pi
1176 adcroft 1.4 }{p}\;\;;\;\;\alpha =\frac{\kappa \Pi \theta }{p}=\frac{\partial \ \Pi }{
1177     \partial p}\theta \;\;;\;\;\frac{D\Pi }{Dt}=\frac{\partial \Pi }{\partial p}
1178 cnh 1.1 \frac{Dp}{Dt}
1179 adcroft 1.4 \end{equation*}
1180 cnh 1.1 where $b=\frac{\partial \ \Pi }{\partial p}\theta $ is the buoyancy.
1181    
1182     The heat equation is obtained by noting that
1183     \begin{equation*}
1184     c_{p}\frac{DT}{Dt}=\frac{D(\Pi \theta )}{Dt}=\Pi \frac{D\theta }{Dt}+\theta
1185 cnh 1.2 \frac{D\Pi }{Dt}=\Pi \frac{D\theta }{Dt}+\alpha \frac{Dp}{Dt}
1186 cnh 1.1 \end{equation*}
1187     and on substituting into (\ref{eq-p-heat-interim}) gives:
1188     \begin{equation}
1189     \Pi \frac{D\theta }{Dt}=\mathcal{Q}
1190     \label{eq:potential-temperature-equation}
1191     \end{equation}
1192     which is in conservative form.
1193    
1194 adcroft 1.4 For convenience in the model we prefer to step forward (\ref
1195 cnh 1.1 {eq:potential-temperature-equation}) rather than (\ref{eq:atmos-heat}).
1196    
1197     \subsubsection{Boundary conditions}
1198    
1199     The upper and lower boundary conditions are :
1200     \begin{eqnarray}
1201     \mbox{at the top:}\;\;p=0 &&\text{, }\omega =\frac{Dp}{Dt}=0 \\
1202     \mbox{at the surface:}\;\;p=p_{s} &&\text{, }\phi =\phi _{topo}=g~Z_{topo}
1203     \label{eq:boundary-condition-atmosphere}
1204     \end{eqnarray}
1205     In $p$-coordinates, the upper boundary acts like a solid boundary ($\omega
1206     =0 $); in $z$-coordinates and the lower boundary is analogous to a free
1207     surface ($\phi $ is imposed and $\omega \neq 0$).
1208    
1209     \subsubsection{Splitting the geo-potential}
1210 jmc 1.22 \label{sec:hpe-p-geo-potential-split}
1211 cnh 1.1
1212     For the purposes of initialization and reducing round-off errors, the model
1213     deals with perturbations from reference (or ``standard'') profiles. For
1214     example, the hydrostatic geopotential associated with the resting atmosphere
1215     is not dynamically relevant and can therefore be subtracted from the
1216     equations. The equations written in terms of perturbations are obtained by
1217     substituting the following definitions into the previous model equations:
1218     \begin{eqnarray}
1219     \theta &=&\theta _{o}+\theta ^{\prime } \label{eq:atmos-ref-prof-theta} \\
1220     \alpha &=&\alpha _{o}+\alpha ^{\prime } \label{eq:atmos-ref-prof-alpha} \\
1221     \phi &=&\phi _{o}+\phi ^{\prime } \label{eq:atmos-ref-prof-phi}
1222     \end{eqnarray}
1223     The reference state (indicated by subscript ``0'') corresponds to
1224     horizontally homogeneous atmosphere at rest ($\theta _{o},\alpha _{o},\phi
1225     _{o}$) with surface pressure $p_{o}(x,y)$ that satisfies $\phi
1226     _{o}(p_{o})=g~Z_{topo}$, defined:
1227     \begin{eqnarray*}
1228     \theta _{o}(p) &=&f^{n}(p) \\
1229     \alpha _{o}(p) &=&\Pi _{p}\theta _{o} \\
1230     \phi _{o}(p) &=&\phi _{topo}-\int_{p_{0}}^{p}\alpha _{o}dp
1231     \end{eqnarray*}
1232     %\begin{eqnarray*}
1233     %\phi'_\alpha & = & \int^p_{p_o} (\alpha_o -\alpha) dp \\
1234     %\phi'_s(x,y,t) & = & \int_{p_o}^{p_s} \alpha dp
1235     %\end{eqnarray*}
1236    
1237     The final form of the HPE's in p coordinates is then:
1238     \begin{eqnarray}
1239 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1240 edhill 1.21 _{h}+\mathbf{\nabla }_{p}\phi ^{\prime } &=&\vec{\mathbf{\mathcal{F}}}
1241     \label{eq:atmos-prime} \\
1242 cnh 1.1 \frac{\partial \phi ^{\prime }}{\partial p}+\alpha ^{\prime } &=&0 \\
1243 adcroft 1.4 \mathbf{\nabla }_{p}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \omega }{
1244 cnh 1.1 \partial p} &=&0 \\
1245     \frac{\partial \Pi }{\partial p}\theta ^{\prime } &=&\alpha ^{\prime } \\
1246 cnh 1.8 \frac{D\theta }{Dt} &=&\frac{\mathcal{Q}}{\Pi }
1247 cnh 1.1 \end{eqnarray}
1248    
1249     \section{Appendix OCEAN}
1250    
1251     \subsection{Equations of motion for the ocean}
1252    
1253     We review here the method by which the standard (Boussinesq, incompressible)
1254     HPE's for the ocean written in z-coordinates are obtained. The
1255     non-Boussinesq equations for oceanic motion are:
1256     \begin{eqnarray}
1257 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1258 cnh 1.1 _{h}+\frac{1}{\rho }\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}} \\
1259     \epsilon _{nh}\frac{Dw}{Dt}+g+\frac{1}{\rho }\frac{\partial p}{\partial z}
1260     &=&\epsilon _{nh}\mathcal{F}_{w} \\
1261 adcroft 1.4 \frac{1}{\rho }\frac{D\rho }{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}
1262 cnh 1.8 _{h}+\frac{\partial w}{\partial z} &=&0 \label{eq-zns-cont}\\
1263     \rho &=&\rho (\theta ,S,p) \label{eq-zns-eos}\\
1264     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-zns-heat}\\
1265     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-zns-salt}
1266     \label{eq:non-boussinesq}
1267 adcroft 1.4 \end{eqnarray}
1268 cnh 1.1 These equations permit acoustics modes, inertia-gravity waves,
1269 cnh 1.10 non-hydrostatic motions, a geostrophic (Rossby) mode and a thermohaline
1270 cnh 1.1 mode. As written, they cannot be integrated forward consistently - if we
1271     step $\rho $ forward in (\ref{eq-zns-cont}), the answer will not be
1272 adcroft 1.4 consistent with that obtained by stepping (\ref{eq-zns-heat}) and (\ref
1273 cnh 1.1 {eq-zns-salt}) and then using (\ref{eq-zns-eos}) to yield $\rho $. It is
1274     therefore necessary to manipulate the system as follows. Differentiating the
1275     EOS (equation of state) gives:
1276    
1277     \begin{equation}
1278     \frac{D\rho }{Dt}=\left. \frac{\partial \rho }{\partial \theta }\right|
1279     _{S,p}\frac{D\theta }{Dt}+\left. \frac{\partial \rho }{\partial S}\right|
1280     _{\theta ,p}\frac{DS}{Dt}+\left. \frac{\partial \rho }{\partial p}\right|
1281     _{\theta ,S}\frac{Dp}{Dt} \label{EOSexpansion}
1282     \end{equation}
1283    
1284 edhill 1.21 Note that $\frac{\partial \rho }{\partial p}=\frac{1}{c_{s}^{2}}$ is
1285     the reciprocal of the sound speed ($c_{s}$) squared. Substituting into
1286     \ref{eq-zns-cont} gives:
1287 cnh 1.1 \begin{equation}
1288 adcroft 1.4 \frac{1}{\rho c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{
1289 cnh 1.1 v}}+\partial _{z}w\approx 0 \label{eq-zns-pressure}
1290     \end{equation}
1291     where we have used an approximation sign to indicate that we have assumed
1292     adiabatic motion, dropping the $\frac{D\theta }{Dt}$ and $\frac{DS}{Dt}$.
1293     Replacing \ref{eq-zns-cont} with \ref{eq-zns-pressure} yields a system that
1294     can be explicitly integrated forward:
1295     \begin{eqnarray}
1296 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1297 cnh 1.1 _{h}+\frac{1}{\rho }\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1298     \label{eq-cns-hmom} \\
1299     \epsilon _{nh}\frac{Dw}{Dt}+g+\frac{1}{\rho }\frac{\partial p}{\partial z}
1300     &=&\epsilon _{nh}\mathcal{F}_{w} \label{eq-cns-hydro} \\
1301 adcroft 1.4 \frac{1}{\rho c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{
1302 cnh 1.1 v}}_{h}+\frac{\partial w}{\partial z} &=&0 \label{eq-cns-cont} \\
1303     \rho &=&\rho (\theta ,S,p) \label{eq-cns-eos} \\
1304     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-cns-heat} \\
1305     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-cns-salt}
1306     \end{eqnarray}
1307    
1308     \subsubsection{Compressible z-coordinate equations}
1309    
1310     Here we linearize the acoustic modes by replacing $\rho $ with $\rho _{o}(z)$
1311     wherever it appears in a product (ie. non-linear term) - this is the
1312     `Boussinesq assumption'. The only term that then retains the full variation
1313     in $\rho $ is the gravitational acceleration:
1314     \begin{eqnarray}
1315 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1316 cnh 1.1 _{h}+\frac{1}{\rho _{o}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1317     \label{eq-zcb-hmom} \\
1318 adcroft 1.4 \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{o}}+\frac{1}{\rho _{o}}
1319 cnh 1.1 \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1320     \label{eq-zcb-hydro} \\
1321 adcroft 1.4 \frac{1}{\rho _{o}c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{
1322 cnh 1.1 \mathbf{v}}_{h}+\frac{\partial w}{\partial z} &=&0 \label{eq-zcb-cont} \\
1323     \rho &=&\rho (\theta ,S,p) \label{eq-zcb-eos} \\
1324     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-zcb-heat} \\
1325     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-zcb-salt}
1326     \end{eqnarray}
1327     These equations still retain acoustic modes. But, because the
1328 adcroft 1.4 ``compressible'' terms are linearized, the pressure equation \ref
1329 cnh 1.1 {eq-zcb-cont} can be integrated implicitly with ease (the time-dependent
1330     term appears as a Helmholtz term in the non-hydrostatic pressure equation).
1331     These are the \emph{truly} compressible Boussinesq equations. Note that the
1332     EOS must have the same pressure dependency as the linearized pressure term,
1333 adcroft 1.4 ie. $\left. \frac{\partial \rho }{\partial p}\right| _{\theta ,S}=\frac{1}{
1334 cnh 1.1 c_{s}^{2}}$, for consistency.
1335    
1336     \subsubsection{`Anelastic' z-coordinate equations}
1337    
1338     The anelastic approximation filters the acoustic mode by removing the
1339 adcroft 1.4 time-dependency in the continuity (now pressure-) equation (\ref{eq-zcb-cont}
1340     ). This could be done simply by noting that $\frac{Dp}{Dt}\approx -g\rho _{o}
1341 cnh 1.1 \frac{Dz}{Dt}=-g\rho _{o}w$, but this leads to an inconsistency between
1342     continuity and EOS. A better solution is to change the dependency on
1343     pressure in the EOS by splitting the pressure into a reference function of
1344     height and a perturbation:
1345     \begin{equation*}
1346 cnh 1.2 \rho =\rho (\theta ,S,p_{o}(z)+\epsilon _{s}p^{\prime })
1347 cnh 1.1 \end{equation*}
1348     Remembering that the term $\frac{Dp}{Dt}$ in continuity comes from
1349     differentiating the EOS, the continuity equation then becomes:
1350     \begin{equation*}
1351 adcroft 1.4 \frac{1}{\rho _{o}c_{s}^{2}}\left( \frac{Dp_{o}}{Dt}+\epsilon _{s}\frac{
1352     Dp^{\prime }}{Dt}\right) +\mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+
1353 cnh 1.2 \frac{\partial w}{\partial z}=0
1354 cnh 1.1 \end{equation*}
1355     If the time- and space-scales of the motions of interest are longer than
1356 adcroft 1.4 those of acoustic modes, then $\frac{Dp^{\prime }}{Dt}<<(\frac{Dp_{o}}{Dt},
1357 cnh 1.1 \mathbf{\nabla }\cdot \vec{\mathbf{v}}_{h})$ in the continuity equations and
1358 adcroft 1.4 $\left. \frac{\partial \rho }{\partial p}\right| _{\theta ,S}\frac{
1359 cnh 1.1 Dp^{\prime }}{Dt}<<\left. \frac{\partial \rho }{\partial p}\right| _{\theta
1360     ,S}\frac{Dp_{o}}{Dt}$ in the EOS (\ref{EOSexpansion}). Thus we set $\epsilon
1361     _{s}=0$, removing the dependency on $p^{\prime }$ in the continuity equation
1362     and EOS. Expanding $\frac{Dp_{o}(z)}{Dt}=-g\rho _{o}w$ then leads to the
1363     anelastic continuity equation:
1364     \begin{equation}
1365 adcroft 1.4 \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z}-
1366 cnh 1.1 \frac{g}{c_{s}^{2}}w=0 \label{eq-za-cont1}
1367     \end{equation}
1368     A slightly different route leads to the quasi-Boussinesq continuity equation
1369 adcroft 1.4 where we use the scaling $\frac{\partial \rho ^{\prime }}{\partial t}+
1370     \mathbf{\nabla }_{3}\cdot \rho ^{\prime }\vec{\mathbf{v}}<<\mathbf{\nabla }
1371 cnh 1.1 _{3}\cdot \rho _{o}\vec{\mathbf{v}}$ yielding:
1372     \begin{equation}
1373 adcroft 1.4 \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{1}{\rho _{o}}\frac{
1374 cnh 1.1 \partial \left( \rho _{o}w\right) }{\partial z}=0 \label{eq-za-cont2}
1375     \end{equation}
1376     Equations \ref{eq-za-cont1} and \ref{eq-za-cont2} are in fact the same
1377     equation if:
1378     \begin{equation}
1379     \frac{1}{\rho _{o}}\frac{\partial \rho _{o}}{\partial z}=\frac{-g}{c_{s}^{2}}
1380     \end{equation}
1381     Again, note that if $\rho _{o}$ is evaluated from prescribed $\theta _{o}$
1382 adcroft 1.4 and $S_{o}$ profiles, then the EOS dependency on $p_{o}$ and the term $\frac{
1383 cnh 1.1 g}{c_{s}^{2}}$ in continuity should be referred to those same profiles. The
1384     full set of `quasi-Boussinesq' or `anelastic' equations for the ocean are
1385     then:
1386     \begin{eqnarray}
1387 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1388 cnh 1.1 _{h}+\frac{1}{\rho _{o}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1389     \label{eq-zab-hmom} \\
1390 adcroft 1.4 \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{o}}+\frac{1}{\rho _{o}}
1391 cnh 1.1 \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1392     \label{eq-zab-hydro} \\
1393 adcroft 1.4 \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{1}{\rho _{o}}\frac{
1394 cnh 1.1 \partial \left( \rho _{o}w\right) }{\partial z} &=&0 \label{eq-zab-cont} \\
1395     \rho &=&\rho (\theta ,S,p_{o}(z)) \label{eq-zab-eos} \\
1396     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-zab-heat} \\
1397     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-zab-salt}
1398     \end{eqnarray}
1399    
1400     \subsubsection{Incompressible z-coordinate equations}
1401    
1402     Here, the objective is to drop the depth dependence of $\rho _{o}$ and so,
1403     technically, to also remove the dependence of $\rho $ on $p_{o}$. This would
1404     yield the ``truly'' incompressible Boussinesq equations:
1405     \begin{eqnarray}
1406 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1407 cnh 1.1 _{h}+\frac{1}{\rho _{c}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}}
1408     \label{eq-ztb-hmom} \\
1409 adcroft 1.4 \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{c}}+\frac{1}{\rho _{c}}
1410 cnh 1.1 \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1411     \label{eq-ztb-hydro} \\
1412     \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z}
1413     &=&0 \label{eq-ztb-cont} \\
1414     \rho &=&\rho (\theta ,S) \label{eq-ztb-eos} \\
1415     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq-ztb-heat} \\
1416     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq-ztb-salt}
1417     \end{eqnarray}
1418     where $\rho _{c}$ is a constant reference density of water.
1419    
1420     \subsubsection{Compressible non-divergent equations}
1421    
1422     The above ``incompressible'' equations are incompressible in both the flow
1423     and the density. In many oceanic applications, however, it is important to
1424     retain compressibility effects in the density. To do this we must split the
1425     density thus:
1426     \begin{equation*}
1427     \rho =\rho _{o}+\rho ^{\prime }
1428 adcroft 1.4 \end{equation*}
1429 cnh 1.1 We then assert that variations with depth of $\rho _{o}$ are unimportant
1430     while the compressible effects in $\rho ^{\prime }$ are:
1431     \begin{equation*}
1432     \rho _{o}=\rho _{c}
1433 adcroft 1.4 \end{equation*}
1434 cnh 1.1 \begin{equation*}
1435     \rho ^{\prime }=\rho (\theta ,S,p_{o}(z))-\rho _{o}
1436 adcroft 1.4 \end{equation*}
1437 cnh 1.1 This then yields what we can call the semi-compressible Boussinesq
1438     equations:
1439     \begin{eqnarray}
1440 adcroft 1.4 \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}}
1441     _{h}+\frac{1}{\rho _{c}}\mathbf{\nabla }_{z}p^{\prime } &=&\vec{\mathbf{
1442 cnh 1.1 \mathcal{F}}} \label{eq:ocean-mom} \\
1443     \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho ^{\prime }}{\rho _{c}}+\frac{1}{\rho
1444     _{c}}\frac{\partial p^{\prime }}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w}
1445     \label{eq:ocean-wmom} \\
1446     \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z}
1447     &=&0 \label{eq:ocean-cont} \\
1448     \rho ^{\prime } &=&\rho (\theta ,S,p_{o}(z))-\rho _{c} \label{eq:ocean-eos}
1449     \\
1450     \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \label{eq:ocean-theta} \\
1451     \frac{DS}{Dt} &=&\mathcal{Q}_{s} \label{eq:ocean-salt}
1452 adcroft 1.4 \end{eqnarray}
1453 cnh 1.1 Note that the hydrostatic pressure of the resting fluid, including that
1454     associated with $\rho _{c}$, is subtracted out since it has no effect on the
1455     dynamics.
1456    
1457     Though necessary, the assumptions that go into these equations are messy
1458     since we essentially assume a different EOS for the reference density and
1459     the perturbation density. Nevertheless, it is the hydrostatic ($\epsilon
1460     _{nh}=0$ form of these equations that are used throughout the ocean modeling
1461     community and referred to as the primitive equations (HPE).
1462    
1463     \section{Appendix:OPERATORS}
1464    
1465     \subsection{Coordinate systems}
1466    
1467     \subsubsection{Spherical coordinates}
1468    
1469     In spherical coordinates, the velocity components in the zonal, meridional
1470     and vertical direction respectively, are given by (see Fig.2) :
1471    
1472     \begin{equation*}
1473 cnh 1.6 u=r\cos \varphi \frac{D\lambda }{Dt}
1474 cnh 1.1 \end{equation*}
1475    
1476     \begin{equation*}
1477 edhill 1.26 v=r\frac{D\varphi }{Dt}
1478 cnh 1.1 \end{equation*}
1479    
1480     \begin{equation*}
1481 cnh 1.2 \dot{r}=\frac{Dr}{Dt}
1482 cnh 1.1 \end{equation*}
1483    
1484 cnh 1.6 Here $\varphi $ is the latitude, $\lambda $ the longitude, $r$ the radial
1485 cnh 1.1 distance of the particle from the center of the earth, $\Omega $ is the
1486     angular speed of rotation of the Earth and $D/Dt$ is the total derivative.
1487    
1488 edhill 1.26 The `grad' ($\nabla $) and `div' ($\nabla\cdot$) operators are defined by, in
1489 cnh 1.1 spherical coordinates:
1490    
1491     \begin{equation*}
1492 cnh 1.6 \nabla \equiv \left( \frac{1}{r\cos \varphi }\frac{\partial }{\partial \lambda }
1493     ,\frac{1}{r}\frac{\partial }{\partial \varphi },\frac{\partial }{\partial r}
1494 cnh 1.2 \right)
1495 cnh 1.1 \end{equation*}
1496    
1497     \begin{equation*}
1498 edhill 1.26 \nabla\cdot v\equiv \frac{1}{r\cos \varphi }\left\{ \frac{\partial u}{\partial
1499 cnh 1.6 \lambda }+\frac{\partial }{\partial \varphi }\left( v\cos \varphi \right) \right\}
1500 cnh 1.2 +\frac{1}{r^{2}}\frac{\partial \left( r^{2}\dot{r}\right) }{\partial r}
1501 cnh 1.1 \end{equation*}
1502    
1503 adcroft 1.4 %tci%\end{document}

  ViewVC Help
Powered by ViewVC 1.1.22