| 1 | % $Header: /u/gcmpack/mitgcmdoc/part1/manual.tex,v 1.3 2001/10/10 16:48:45 cnh Exp $ | 
| 2 | % $Name:  $ | 
| 3 |  | 
| 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 | \part{MIT GCM basics} | 
| 36 |  | 
| 37 | % Section: Overview | 
| 38 |  | 
| 39 | % $Header: /u/gcmpack/mitgcmdoc/part1/manual.tex,v 1.3 2001/10/10 16:48:45 cnh Exp $ | 
| 40 | % $Name:  $ | 
| 41 |  | 
| 42 | \section{Introduction} | 
| 43 |  | 
| 44 | This documentation provides the reader with the information necessary to | 
| 45 | carry out numerical experiments using MITgcm. It gives a comprehensive | 
| 46 | description of the continuous equations on which the model is based, the | 
| 47 | numerical algorithms the model employs and a description of the associated | 
| 48 | program code. Along with the hydrodynamical kernel, physical and | 
| 49 | biogeochemical parameterizations of key atmospheric and oceanic processes | 
| 50 | are available. A number of examples illustrating the use of the model in | 
| 51 | both process and general circulation studies of the atmosphere and ocean are | 
| 52 | also presented. | 
| 53 |  | 
| 54 | MITgcm has a number of novel aspects: | 
| 55 |  | 
| 56 | \begin{itemize} | 
| 57 | \item it can be used to study both atmospheric and oceanic phenomena; one | 
| 58 | hydrodynamical kernel is used to drive forward both atmospheric and oceanic | 
| 59 | models - see fig | 
| 60 | \marginpar{ | 
| 61 | Fig.1 One model}\ref{fig:onemodel} | 
| 62 |  | 
| 63 | %% CNHbegin | 
| 64 | %notci%\input{part1/one_model_figure} | 
| 65 | %% CNHend | 
| 66 |  | 
| 67 | \item it has a non-hydrostatic capability and so can be used to study both | 
| 68 | small-scale and large scale processes - see fig | 
| 69 | \marginpar{ | 
| 70 | Fig.2 All scales}\ref{fig:all-scales} | 
| 71 |  | 
| 72 | %% CNHbegin | 
| 73 | %notci%\input{part1/all_scales_figure} | 
| 74 | %% CNHend | 
| 75 |  | 
| 76 | \item finite volume techniques are employed yielding an intuitive | 
| 77 | discretization and support for the treatment of irregular geometries using | 
| 78 | orthogonal curvilinear grids and shaved cells - see fig | 
| 79 | \marginpar{ | 
| 80 | Fig.3 Finite volumes}\ref{fig:finite-volumes} | 
| 81 |  | 
| 82 | %% CNHbegin | 
| 83 | %notci%\input{part1/fvol_figure} | 
| 84 | %% CNHend | 
| 85 |  | 
| 86 | \item tangent linear and adjoint counterparts are automatically maintained | 
| 87 | along with the forward model, permitting sensitivity and optimization | 
| 88 | studies. | 
| 89 |  | 
| 90 | \item the model is developed to perform efficiently on a wide variety of | 
| 91 | computational platforms. | 
| 92 | \end{itemize} | 
| 93 |  | 
| 94 | Key publications reporting on and charting the development of the model are | 
| 95 | listed in an Appendix. | 
| 96 |  | 
| 97 | We begin by briefly showing some of the results of the model in action to | 
| 98 | give a feel for the wide range of problems that can be addressed using it. | 
| 99 | \pagebreak | 
| 100 |  | 
| 101 | % $Header: /u/gcmpack/mitgcmdoc/part1/manual.tex,v 1.3 2001/10/10 16:48:45 cnh Exp $ | 
| 102 | % $Name:  $ | 
| 103 |  | 
| 104 | \section{Illustrations of the model in action} | 
| 105 |  | 
| 106 | The MITgcm has been designed and used to model a wide range of phenomena, | 
| 107 | from convection on the scale of meters in the ocean to the global pattern of | 
| 108 | atmospheric winds - see fig.2\ref{fig:all-scales}. To give a flavor of the | 
| 109 | kinds of problems the model has been used to study, we briefly describe some | 
| 110 | of them here. A more detailed description of the underlying formulation, | 
| 111 | numerical algorithm and implementation that lie behind these calculations is | 
| 112 | given later. Indeed many of the illustrative examples shown below can be | 
| 113 | easily reproduced: simply download the model (the minimum you need is a PC | 
| 114 | running linux, together with a FORTRAN\ 77 compiler) and follow the examples | 
| 115 | described in detail in the documentation. | 
| 116 |  | 
| 117 | \subsection{Global atmosphere: `Held-Suarez' benchmark} | 
| 118 |  | 
| 119 | A novel feature of MITgcm is its ability to simulate both atmospheric and | 
| 120 | oceanographic flows at both small and large scales. | 
| 121 |  | 
| 122 | Fig.E1a.\ref{fig:eddy_cs} shows an instantaneous plot of the 500$mb$ | 
| 123 | temperature field obtained using the atmospheric isomorph of MITgcm run at | 
| 124 | 2.8$^{\circ }$ resolution on the cubed sphere. We see cold air over the pole | 
| 125 | (blue) and warm air along an equatorial band (red). Fully developed | 
| 126 | baroclinic eddies spawned in the northern hemisphere storm track are | 
| 127 | evident. There are no mountains or land-sea contrast in this calculation, | 
| 128 | but you can easily put them in. The model is driven by relaxation to a | 
| 129 | radiative-convective equilibrium profile, following the description set out | 
| 130 | in Held and Suarez; 1994 designed to test atmospheric hydrodynamical cores - | 
| 131 | there are no mountains or land-sea contrast. | 
| 132 |  | 
| 133 | %% CNHbegin | 
| 134 | %notci%\input{part1/cubic_eddies_figure} | 
| 135 | %% CNHend | 
| 136 |  | 
| 137 | As described in Adcroft (2001), a `cubed sphere' is used to discretize the | 
| 138 | globe permitting a uniform gridding and obviated the need to fourier filter. | 
| 139 | The `vector-invariant' form of MITgcm supports any orthogonal curvilinear | 
| 140 | grid, of which the cubed sphere is just one of many choices. | 
| 141 |  | 
| 142 | Fig.E1b shows the 5-year mean, zonally averaged potential temperature, zonal | 
| 143 | wind and meridional overturning streamfunction from a 20-level version of | 
| 144 | the model. It compares favorable with more conventional spatial | 
| 145 | discretization approaches. | 
| 146 |  | 
| 147 | A regular spherical lat-lon grid can also be used. | 
| 148 |  | 
| 149 | %% CNHbegin | 
| 150 | %notci%\input{part1/hs_zave_u_figure} | 
| 151 | %% CNHend | 
| 152 |  | 
| 153 | \subsection{Ocean gyres} | 
| 154 |  | 
| 155 | Baroclinic instability is a ubiquitous process in the ocean, as well as the | 
| 156 | atmosphere. Ocean eddies play an important role in modifying the | 
| 157 | hydrographic structure and current systems of the oceans. Coarse resolution | 
| 158 | models of the oceans cannot resolve the eddy field and yield rather broad, | 
| 159 | diffusive patterns of ocean currents. But if the resolution of our models is | 
| 160 | increased until the baroclinic instability process is resolved, numerical | 
| 161 | solutions of a different and much more realistic kind, can be obtained. | 
| 162 |  | 
| 163 | Fig. ?.? shows the surface temperature and velocity field obtained from | 
| 164 | MITgcm run at $\frac{1}{6}^{\circ }$ horizontal resolution on a $lat-lon$ | 
| 165 | grid in which the pole has been rotated by 90$^{\circ }$ on to the equator | 
| 166 | (to avoid the converging of meridian in northern latitudes). 21 vertical | 
| 167 | levels are used in the vertical with a `lopped cell' representation of | 
| 168 | topography. The development and propagation of anomalously warm and cold | 
| 169 | eddies can be clearly been seen in the Gulf Stream region. The transport of | 
| 170 | warm water northward by the mean flow of the Gulf Stream is also clearly | 
| 171 | visible. | 
| 172 |  | 
| 173 | %% CNHbegin | 
| 174 | %notci%\input{part1/ocean_gyres_figure} | 
| 175 | %% CNHend | 
| 176 |  | 
| 177 |  | 
| 178 | \subsection{Global ocean circulation} | 
| 179 |  | 
| 180 | Fig.E2a shows the pattern of ocean currents at the surface of a 4$^{\circ }$ | 
| 181 | global ocean model run with 15 vertical levels. Lopped cells are used to | 
| 182 | represent topography on a regular $lat-lon$ grid extending from 70$^{\circ | 
| 183 | }N $ to 70$^{\circ }S$. The model is driven using monthly-mean winds with | 
| 184 | mixed boundary conditions on temperature and salinity at the surface. The | 
| 185 | transfer properties of ocean eddies, convection and mixing is parameterized | 
| 186 | in this model. | 
| 187 |  | 
| 188 | Fig.E2b shows the meridional overturning circulation of the global ocean in | 
| 189 | Sverdrups. | 
| 190 |  | 
| 191 | %%CNHbegin | 
| 192 | %notci%\input{part1/global_circ_figure} | 
| 193 | %%CNHend | 
| 194 |  | 
| 195 | \subsection{Convection and mixing over topography} | 
| 196 |  | 
| 197 | Dense plumes generated by localized cooling on the continental shelf of the | 
| 198 | ocean may be influenced by rotation when the deformation radius is smaller | 
| 199 | than the width of the cooling region. Rather than gravity plumes, the | 
| 200 | mechanism for moving dense fluid down the shelf is then through geostrophic | 
| 201 | eddies. The simulation shown in the figure (blue is cold dense fluid, red is | 
| 202 | warmer, lighter fluid) employs the non-hydrostatic capability of MITgcm to | 
| 203 | trigger convection by surface cooling. The cold, dense water falls down the | 
| 204 | slope but is deflected along the slope by rotation. It is found that | 
| 205 | entrainment in the vertical plane is reduced when rotational control is | 
| 206 | strong, and replaced by lateral entrainment due to the baroclinic | 
| 207 | instability of the along-slope current. | 
| 208 |  | 
| 209 | %%CNHbegin | 
| 210 | %notci%\input{part1/convect_and_topo} | 
| 211 | %%CNHend | 
| 212 |  | 
| 213 | \subsection{Boundary forced internal waves} | 
| 214 |  | 
| 215 | The unique ability of MITgcm to treat non-hydrostatic dynamics in the | 
| 216 | presence of complex geometry makes it an ideal tool to study internal wave | 
| 217 | dynamics and mixing in oceanic canyons and ridges driven by large amplitude | 
| 218 | barotropic tidal currents imposed through open boundary conditions. | 
| 219 |  | 
| 220 | Fig. ?.? shows the influence of cross-slope topographic variations on | 
| 221 | internal wave breaking - the cross-slope velocity is in color, the density | 
| 222 | contoured. The internal waves are excited by application of open boundary | 
| 223 | conditions on the left.\ They propagate to the sloping boundary (represented | 
| 224 | using MITgcm's finite volume spatial discretization) where they break under | 
| 225 | nonhydrostatic dynamics. | 
| 226 |  | 
| 227 | %%CNHbegin | 
| 228 | %notci%\input{part1/boundary_forced_waves} | 
| 229 | %%CNHend | 
| 230 |  | 
| 231 | \subsection{Parameter sensitivity using the adjoint of MITgcm} | 
| 232 |  | 
| 233 | Forward and tangent linear counterparts of MITgcm are supported using an | 
| 234 | `automatic adjoint compiler'. These can be used in parameter sensitivity and | 
| 235 | data assimilation studies. | 
| 236 |  | 
| 237 | As one example of application of the MITgcm adjoint, Fig.E4 maps the | 
| 238 | gradient $\frac{\partial J}{\partial \mathcal{H}}$where $J$ is the magnitude | 
| 239 | of the overturning streamfunction shown in fig?.? at 40$^{\circ }$N and $ | 
| 240 | \mathcal{H}$ is the air-sea heat flux 100 years before. We see that $J$ is | 
| 241 | sensitive to heat fluxes over the Labrador Sea, one of the important sources | 
| 242 | of deep water for the thermohaline circulations. This calculation also | 
| 243 | yields sensitivities to all other model parameters. | 
| 244 |  | 
| 245 | %%CNHbegin | 
| 246 | %notci%\input{part1/adj_hf_ocean_figure} | 
| 247 | %%CNHend | 
| 248 |  | 
| 249 | \subsection{Global state estimation of the ocean} | 
| 250 |  | 
| 251 | An important application of MITgcm is in state estimation of the global | 
| 252 | ocean circulation. An appropriately defined `cost function', which measures | 
| 253 | the departure of the model from observations (both remotely sensed and | 
| 254 | insitu) over an interval of time, is minimized by adjusting `control | 
| 255 | parameters' such as air-sea fluxes, the wind field, the initial conditions | 
| 256 | etc. Figure ?.? shows an estimate of the time-mean surface elevation of the | 
| 257 | ocean obtained by bringing the model in to consistency with altimetric and | 
| 258 | in-situ observations over the period 1992-1997. | 
| 259 |  | 
| 260 | %% CNHbegin | 
| 261 | %notci%\input{part1/globes_figure} | 
| 262 | %% CNHend | 
| 263 |  | 
| 264 | \subsection{Ocean biogeochemical cycles} | 
| 265 |  | 
| 266 | MITgcm is being used to study global biogeochemical cycles in the ocean. For | 
| 267 | example one can study the effects of interannual changes in meteorological | 
| 268 | forcing and upper ocean circulation on the fluxes of carbon dioxide and | 
| 269 | oxygen between the ocean and atmosphere. The figure shows the annual air-sea | 
| 270 | flux of oxygen and its relation to density outcrops in the southern oceans | 
| 271 | from a single year of a global, interannually varying simulation. | 
| 272 |  | 
| 273 | %%CNHbegin | 
| 274 | %notci%\input{part1/biogeo_figure} | 
| 275 | %%CNHend | 
| 276 |  | 
| 277 | \subsection{Simulations of laboratory experiments} | 
| 278 |  | 
| 279 | Figure ?.? shows MITgcm being used to simulate a laboratory experiment | 
| 280 | enquiring in to the dynamics of the Antarctic Circumpolar Current (ACC). An | 
| 281 | initially homogeneous tank of water ($1m$ in diameter) is driven from its | 
| 282 | free surface by a rotating heated disk. The combined action of mechanical | 
| 283 | and thermal forcing creates a lens of fluid which becomes baroclinically | 
| 284 | unstable. The stratification and depth of penetration of the lens is | 
| 285 | arrested by its instability in a process analogous to that whic sets the | 
| 286 | stratification of the ACC. | 
| 287 |  | 
| 288 | %%CNHbegin | 
| 289 | %notci%\input{part1/lab_figure} | 
| 290 | %%CNHend | 
| 291 |  | 
| 292 | % $Header: /u/gcmpack/mitgcmdoc/part1/manual.tex,v 1.3 2001/10/10 16:48:45 cnh Exp $ | 
| 293 | % $Name:  $ | 
| 294 |  | 
| 295 | \section{Continuous equations in `r' coordinates} | 
| 296 |  | 
| 297 | To render atmosphere and ocean models from one dynamical core we exploit | 
| 298 | `isomorphisms' between equation sets that govern the evolution of the | 
| 299 | respective fluids - see fig.4 | 
| 300 | \marginpar{ | 
| 301 | Fig.4. Isomorphisms}. One system of hydrodynamical equations is written down | 
| 302 | and encoded. The model variables have different interpretations depending on | 
| 303 | whether the atmosphere or ocean is being studied. Thus, for example, the | 
| 304 | vertical coordinate `$r$' is interpreted as pressure, $p$, if we are | 
| 305 | modeling the atmosphere and height, $z$, if we are modeling the ocean. | 
| 306 |  | 
| 307 | %%CNHbegin | 
| 308 | %notci%\input{part1/zandpcoord_figure.tex} | 
| 309 | %%CNHend | 
| 310 |  | 
| 311 | The state of the fluid at any time is characterized by the distribution of | 
| 312 | velocity $\vec{\mathbf{v}}$, active tracers $\theta $ and $S$, a | 
| 313 | `geopotential' $\phi $ and density $\rho =\rho (\theta ,S,p)$ which may | 
| 314 | depend on $\theta $, $S$, and $p$. The equations that govern the evolution | 
| 315 | of these fields, obtained by applying the laws of classical mechanics and | 
| 316 | thermodynamics to a Boussinesq, Navier-Stokes fluid are, written in terms of | 
| 317 | a generic vertical coordinate, $r$, see fig.5 | 
| 318 | \marginpar{ | 
| 319 | Fig.5 The vertical coordinate of model}: | 
| 320 |  | 
| 321 | %%CNHbegin | 
| 322 | %notci%\input{part1/vertcoord_figure.tex} | 
| 323 | %%CNHend | 
| 324 |  | 
| 325 | \begin{equation*} | 
| 326 | \frac{D\vec{\mathbf{v}_{h}}}{Dt}+\left( 2\vec{\Omega}\times \vec{\mathbf{v}} | 
| 327 | \right) _{h}+\mathbf{\nabla }_{h}\phi =\mathcal{F}_{\vec{\mathbf{v}_{h}}} | 
| 328 | \text{ horizontal mtm} | 
| 329 | \end{equation*} | 
| 330 |  | 
| 331 | \begin{equation*} | 
| 332 | \frac{D\dot{r}}{Dt}+\widehat{k}\cdot \left( 2\vec{\Omega}\times \vec{\mathbf{ | 
| 333 | v}}\right) +\frac{\partial \phi }{\partial r}+b=\mathcal{F}_{\dot{r}}\text{ | 
| 334 | vertical mtm} | 
| 335 | \end{equation*} | 
| 336 |  | 
| 337 | \begin{equation} | 
| 338 | \mathbf{\nabla }_{h}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \dot{r}}{ | 
| 339 | \partial r}=0\text{ continuity}  \label{eq:continuous} | 
| 340 | \end{equation} | 
| 341 |  | 
| 342 | \begin{equation*} | 
| 343 | b=b(\theta ,S,r)\text{ equation of state} | 
| 344 | \end{equation*} | 
| 345 |  | 
| 346 | \begin{equation*} | 
| 347 | \frac{D\theta }{Dt}=\mathcal{Q}_{\theta }\text{ potential temperature} | 
| 348 | \end{equation*} | 
| 349 |  | 
| 350 | \begin{equation*} | 
| 351 | \frac{DS}{Dt}=\mathcal{Q}_{S}\text{ humidity/salinity} | 
| 352 | \end{equation*} | 
| 353 |  | 
| 354 | Here: | 
| 355 |  | 
| 356 | \begin{equation*} | 
| 357 | r\text{ is the vertical coordinate} | 
| 358 | \end{equation*} | 
| 359 |  | 
| 360 | \begin{equation*} | 
| 361 | \frac{D}{Dt}=\frac{\partial }{\partial t}+\vec{\mathbf{v}}\cdot \nabla \text{ | 
| 362 | is the total derivative} | 
| 363 | \end{equation*} | 
| 364 |  | 
| 365 | \begin{equation*} | 
| 366 | \mathbf{\nabla }=\mathbf{\nabla }_{h}+\widehat{k}\frac{\partial }{\partial r} | 
| 367 | \text{ is the `grad' operator} | 
| 368 | \end{equation*} | 
| 369 | with $\mathbf{\nabla }_{h}$ operating in the horizontal and $\widehat{k} | 
| 370 | \frac{\partial }{\partial r}$ operating in the vertical, where $\widehat{k}$ | 
| 371 | is a unit vector in the vertical | 
| 372 |  | 
| 373 | \begin{equation*} | 
| 374 | t\text{ is time} | 
| 375 | \end{equation*} | 
| 376 |  | 
| 377 | \begin{equation*} | 
| 378 | \vec{\mathbf{v}}=(u,v,\dot{r})=(\vec{\mathbf{v}}_{h},\dot{r})\text{ is the | 
| 379 | velocity} | 
| 380 | \end{equation*} | 
| 381 |  | 
| 382 | \begin{equation*} | 
| 383 | \phi \text{ is the `pressure'/`geopotential'} | 
| 384 | \end{equation*} | 
| 385 |  | 
| 386 | \begin{equation*} | 
| 387 | \vec{\Omega}\text{ is the Earth's rotation} | 
| 388 | \end{equation*} | 
| 389 |  | 
| 390 | \begin{equation*} | 
| 391 | b\text{ is the `buoyancy'} | 
| 392 | \end{equation*} | 
| 393 |  | 
| 394 | \begin{equation*} | 
| 395 | \theta \text{ is potential temperature} | 
| 396 | \end{equation*} | 
| 397 |  | 
| 398 | \begin{equation*} | 
| 399 | S\text{ is specific humidity in the atmosphere; salinity in the ocean} | 
| 400 | \end{equation*} | 
| 401 |  | 
| 402 | \begin{equation*} | 
| 403 | \mathcal{F}_{\vec{\mathbf{v}}}\text{ are forcing and dissipation of }\vec{ | 
| 404 | \mathbf{v}} | 
| 405 | \end{equation*} | 
| 406 |  | 
| 407 | \begin{equation*} | 
| 408 | \mathcal{Q}_{\theta }\mathcal{\ }\text{are forcing and dissipation of }\theta | 
| 409 | \end{equation*} | 
| 410 |  | 
| 411 | \begin{equation*} | 
| 412 | \mathcal{Q}_{S}\mathcal{\ }\text{are forcing and dissipation of }S | 
| 413 | \end{equation*} | 
| 414 |  | 
| 415 | The $\mathcal{F}^{\prime }s$ and $\mathcal{Q}^{\prime }s$ are provided by | 
| 416 | extensive `physics' packages for atmosphere and ocean described in Chapter 6. | 
| 417 |  | 
| 418 | \subsection{Kinematic Boundary conditions} | 
| 419 |  | 
| 420 | \subsubsection{vertical} | 
| 421 |  | 
| 422 | at fixed and moving $r$ surfaces we set (see fig.5): | 
| 423 |  | 
| 424 | \begin{equation} | 
| 425 | \dot{r}=0atr=R_{fixed}(x,y)\text{ (ocean bottom, top of the atmosphere)} | 
| 426 | \label{eq:fixedbc} | 
| 427 | \end{equation} | 
| 428 |  | 
| 429 | \begin{equation} | 
| 430 | \dot{r}=\frac{Dr}{Dt}atr=R_{moving}\text{ \ | 
| 431 | (oceansurface,bottomoftheatmosphere)}  \label{eq:movingbc} | 
| 432 | \end{equation} | 
| 433 |  | 
| 434 | Here | 
| 435 |  | 
| 436 | \begin{equation*} | 
| 437 | R_{moving}=R_{o}+\eta | 
| 438 | \end{equation*} | 
| 439 | where $R_{o}(x,y)$ is the `$r-$value' (height or pressure, depending on | 
| 440 | whether we are in the atmosphere or ocean) of the `moving surface' in the | 
| 441 | resting fluid and $\eta $ is the departure from $R_{o}(x,y)$ in the presence | 
| 442 | of motion. | 
| 443 |  | 
| 444 | \subsubsection{horizontal} | 
| 445 |  | 
| 446 | \begin{equation} | 
| 447 | \vec{\mathbf{v}}\cdot \vec{\mathbf{n}}=0  \label{eq:noflow} | 
| 448 | \end{equation} | 
| 449 | where $\vec{\mathbf{n}}$ is the normal to a solid boundary. | 
| 450 |  | 
| 451 | \subsection{Atmosphere} | 
| 452 |  | 
| 453 | In the atmosphere, see fig.5, we interpret: | 
| 454 |  | 
| 455 | \begin{equation} | 
| 456 | r=p\text{ is the pressure}  \label{eq:atmos-r} | 
| 457 | \end{equation} | 
| 458 |  | 
| 459 | \begin{equation} | 
| 460 | \dot{r}=\frac{Dp}{Dt}=\omega \text{ is the vertical velocity in }p\text{ | 
| 461 | coordinates}  \label{eq:atmos-omega} | 
| 462 | \end{equation} | 
| 463 |  | 
| 464 | \begin{equation} | 
| 465 | \phi =g\,z\text{ is the geopotential height}  \label{eq:atmos-phi} | 
| 466 | \end{equation} | 
| 467 |  | 
| 468 | \begin{equation} | 
| 469 | b=\frac{\partial \Pi }{\partial p}\theta \text{ is the buoyancy} | 
| 470 | \label{eq:atmos-b} | 
| 471 | \end{equation} | 
| 472 |  | 
| 473 | \begin{equation} | 
| 474 | \theta =T(\frac{p_{c}}{p})^{\kappa }\text{ is potential temperature} | 
| 475 | \label{eq:atmos-theta} | 
| 476 | \end{equation} | 
| 477 |  | 
| 478 | \begin{equation} | 
| 479 | S=q,\text{ is the specific humidity}  \label{eq:atmos-s} | 
| 480 | \end{equation} | 
| 481 | where | 
| 482 |  | 
| 483 | \begin{equation*} | 
| 484 | T\text{ is absolute temperature} | 
| 485 | \end{equation*} | 
| 486 | \begin{equation*} | 
| 487 | p\text{ is the pressure} | 
| 488 | \end{equation*} | 
| 489 | \begin{eqnarray*} | 
| 490 | &&z\text{ is the height of the pressure surface} \\ | 
| 491 | &&g\text{ is the acceleration due to gravity} | 
| 492 | \end{eqnarray*} | 
| 493 |  | 
| 494 | In the above the ideal gas law, $p=\rho RT$, has been expressed in terms of | 
| 495 | the Exner function $\Pi (p)$ given by (see Appendix Atmosphere) | 
| 496 | \begin{equation} | 
| 497 | \Pi (p)=c_{p}(\frac{p}{p_{c}})^{\kappa }  \label{eq:exner} | 
| 498 | \end{equation} | 
| 499 | where $p_{c}$ is a reference pressure and $\kappa =R/c_{p}$ with $R$ the gas | 
| 500 | constant and $c_{p}$ the specific heat of air at constant pressure. | 
| 501 |  | 
| 502 | At the top of the atmosphere (which is `fixed' in our $r$ coordinate): | 
| 503 |  | 
| 504 | \begin{equation*} | 
| 505 | R_{fixed}=p_{top}=0 | 
| 506 | \end{equation*} | 
| 507 | In a resting atmosphere the elevation of the mountains at the bottom is | 
| 508 | given by | 
| 509 | \begin{equation*} | 
| 510 | R_{moving}=R_{o}(x,y)=p_{o}(x,y) | 
| 511 | \end{equation*} | 
| 512 | i.e. the (hydrostatic) pressure at the top of the mountains in a resting | 
| 513 | atmosphere. | 
| 514 |  | 
| 515 | The boundary conditions at top and bottom are given by: | 
| 516 |  | 
| 517 | \begin{eqnarray} | 
| 518 | &&\omega =0~\text{at }r=R_{fixed} \text{ (top of the atmosphere)} | 
| 519 | \label{eq:fixed-bc-atmos} \\ | 
| 520 | \omega &=&\frac{Dp_{s}}{Dt}\text{; at }r=R_{moving}\text{ (bottom of the | 
| 521 | atmosphere)}  \label{eq:moving-bc-atmos} | 
| 522 | \end{eqnarray} | 
| 523 |  | 
| 524 | Then the (hydrostatic form of) eq(\ref{eq:continuous}) yields a consistent | 
| 525 | set of atmospheric equations which, for convenience, are written out in $p$ | 
| 526 | coordinates in Appendix Atmosphere - see eqs(\ref{eq:atmos-prime}). | 
| 527 |  | 
| 528 | \subsection{Ocean} | 
| 529 |  | 
| 530 | In the ocean we interpret: | 
| 531 | \begin{eqnarray} | 
| 532 | r &=&z\text{ is the height}  \label{eq:ocean-z} \\ | 
| 533 | \dot{r} &=&\frac{Dz}{Dt}=w\text{ is the vertical velocity} | 
| 534 | \label{eq:ocean-w} \\ | 
| 535 | \phi &=&\frac{p}{\rho _{c}}\text{ is the pressure}  \label{eq:ocean-p} \\ | 
| 536 | b(\theta ,S,r) &=&\frac{g}{\rho _{c}}\left( \rho (\theta ,S,r)-\rho | 
| 537 | _{c}\right) \text{ is the buoyancy}  \label{eq:ocean-b} | 
| 538 | \end{eqnarray} | 
| 539 | where $\rho _{c}$ is a fixed reference density of water and $g$ is the | 
| 540 | acceleration due to gravity.\noindent | 
| 541 |  | 
| 542 | In the above | 
| 543 |  | 
| 544 | At the bottom of the ocean: $R_{fixed}(x,y)=-H(x,y)$. | 
| 545 |  | 
| 546 | The surface of the ocean is given by: $R_{moving}=\eta $ | 
| 547 |  | 
| 548 | The position of the resting free surface of the ocean is given by $ | 
| 549 | R_{o}=Z_{o}=0$. | 
| 550 |  | 
| 551 | Boundary conditions are: | 
| 552 |  | 
| 553 | \begin{eqnarray} | 
| 554 | w &=&0~\text{at }r=R_{fixed}\text{ (ocean bottom)}  \label{eq:fixed-bc-ocean} | 
| 555 | \\ | 
| 556 | w &=&\frac{D\eta }{Dt}\text{ at }r=R_{moving}=\eta \text{ (ocean surface) | 
| 557 | \label{eq:moving-bc-ocean}} | 
| 558 | \end{eqnarray} | 
| 559 | where $\eta $ is the elevation of the free surface. | 
| 560 |  | 
| 561 | Then eq(\ref{eq:continuous}) yields a consistent set of oceanic equations | 
| 562 | which, for convenience, are written out in $z$ coordinates in Appendix Ocean | 
| 563 | - see eqs(\ref{eq:ocean-mom}) to (\ref{eq:ocean-salt}). | 
| 564 |  | 
| 565 | \subsection{Hydrostatic, Quasi-hydrostatic, Quasi-nonhydrostatic and | 
| 566 | Non-hydrostatic forms} | 
| 567 |  | 
| 568 | Let us separate $\phi $ in to surface, hydrostatic and non-hydrostatic terms: | 
| 569 |  | 
| 570 | \begin{equation} | 
| 571 | \phi (x,y,r)=\phi _{s}(x,y)+\phi _{hyd}(x,y,r)+\phi _{nh}(x,y,r) | 
| 572 | \label{eq:phi-split} | 
| 573 | \end{equation} | 
| 574 | and write eq(\ref{incompressible}a,b) in the form: | 
| 575 |  | 
| 576 | \begin{equation} | 
| 577 | \frac{\partial \vec{\mathbf{v}_{h}}}{\partial t}+\mathbf{\nabla }_{h}\phi | 
| 578 | _{s}+\mathbf{\nabla }_{h}\phi _{hyd}+\epsilon _{nh}\mathbf{\nabla }_{h}\phi | 
| 579 | _{nh}=\vec{\mathbf{G}}_{\vec{v}_{h}}  \label{eq:mom-h} | 
| 580 | \end{equation} | 
| 581 |  | 
| 582 | \begin{equation} | 
| 583 | \frac{\partial \phi _{hyd}}{\partial r}=-b  \label{eq:hydrostatic} | 
| 584 | \end{equation} | 
| 585 |  | 
| 586 | \begin{equation} | 
| 587 | \epsilon _{nh}\frac{\partial \dot{r}}{\partial t}+\frac{\partial \phi _{nh}}{ | 
| 588 | \partial r}=G_{\dot{r}}  \label{eq:mom-w} | 
| 589 | \end{equation} | 
| 590 | Here $\epsilon _{nh}$ is a non-hydrostatic parameter. | 
| 591 |  | 
| 592 | The $\left( \vec{\mathbf{G}}_{\vec{v}},G_{\dot{r}}\right) $ in eq(\ref | 
| 593 | {eq:mom-h}) and (\ref{eq:mom-w}) represent advective, metric and Coriolis | 
| 594 | terms in the momentum equations. In spherical coordinates they take the form | 
| 595 | \footnote{ | 
| 596 | In the hydrostatic primitive equations (\textbf{HPE}) all underlined terms | 
| 597 | in (\ref{eq:gu-speherical}), (\ref{eq:gv-spherical}) and (\ref | 
| 598 | {eq:gw-spherical}) are omitted; the singly-underlined terms are included in | 
| 599 | the quasi-hydrostatic model (\textbf{QH}). The fully non-hydrostatic model ( | 
| 600 | \textbf{NH}) includes all terms.} - see Marshall et al 1997a for a full | 
| 601 | discussion: | 
| 602 |  | 
| 603 | \begin{equation} | 
| 604 | \left. | 
| 605 | \begin{tabular}{l} | 
| 606 | $G_{u}=-\vec{\mathbf{v}}.\nabla u$ \\ | 
| 607 | $-\left\{ \underline{\frac{u\dot{r}}{{r}}}-\frac{uv\tan lat}{{r}}\right\} $ | 
| 608 | \\ | 
| 609 | $-\left\{ -2\Omega v\sin lat+\underline{2\Omega \dot{r}\cos lat}\right\} $ | 
| 610 | \\ | 
| 611 | $+\mathcal{F}_{u}$ | 
| 612 | \end{tabular} | 
| 613 | \ \right\} \left\{ | 
| 614 | \begin{tabular}{l} | 
| 615 | \textit{advection} \\ | 
| 616 | \textit{metric} \\ | 
| 617 | \textit{Coriolis} \\ | 
| 618 | \textit{\ Forcing/Dissipation} | 
| 619 | \end{tabular} | 
| 620 | \ \right. \qquad  \label{eq:gu-speherical} | 
| 621 | \end{equation} | 
| 622 |  | 
| 623 | \begin{equation} | 
| 624 | \left. | 
| 625 | \begin{tabular}{l} | 
| 626 | $G_{v}=-\vec{\mathbf{v}}.\nabla v$ \\ | 
| 627 | $-\left\{ \underline{\frac{v\dot{r}}{{r}}}-\frac{u^{2}\tan lat}{{r}}\right\} | 
| 628 | $ \\ | 
| 629 | $-\left\{ -2\Omega u\sin lat\right\} $ \\ | 
| 630 | $+\mathcal{F}_{v}$ | 
| 631 | \end{tabular} | 
| 632 | \ \right\} \left\{ | 
| 633 | \begin{tabular}{l} | 
| 634 | \textit{advection} \\ | 
| 635 | \textit{metric} \\ | 
| 636 | \textit{Coriolis} \\ | 
| 637 | \textit{\ Forcing/Dissipation} | 
| 638 | \end{tabular} | 
| 639 | \ \right. \qquad  \label{eq:gv-spherical} | 
| 640 | \end{equation} | 
| 641 | \qquad \qquad \qquad \qquad \qquad | 
| 642 |  | 
| 643 | \begin{equation} | 
| 644 | \left. | 
| 645 | \begin{tabular}{l} | 
| 646 | $G_{\dot{r}}=-\underline{\underline{\vec{\mathbf{v}}.\nabla \dot{r}}}$ \\ | 
| 647 | $+\left\{ \underline{\frac{u^{_{^{2}}}+v^{2}}{{r}}}\right\} $ \\ | 
| 648 | ${+}\underline{{2\Omega u\cos lat}}$ \\ | 
| 649 | $\underline{\underline{\mathcal{F}_{\dot{r}}}}$ | 
| 650 | \end{tabular} | 
| 651 | \ \right\} \left\{ | 
| 652 | \begin{tabular}{l} | 
| 653 | \textit{advection} \\ | 
| 654 | \textit{metric} \\ | 
| 655 | \textit{Coriolis} \\ | 
| 656 | \textit{\ Forcing/Dissipation} | 
| 657 | \end{tabular} | 
| 658 | \ \right.  \label{eq:gw-spherical} | 
| 659 | \end{equation} | 
| 660 | \qquad \qquad \qquad \qquad \qquad | 
| 661 |  | 
| 662 | In the above `${r}$' is the distance from the center of the earth and `$lat$ | 
| 663 | ' is latitude. | 
| 664 |  | 
| 665 | Grad and div operators in spherical coordinates are defined in appendix | 
| 666 | OPERATORS. | 
| 667 | \marginpar{ | 
| 668 | Fig.6 Spherical polar coordinate system.} | 
| 669 |  | 
| 670 | %%CNHbegin | 
| 671 | %notci%\input{part1/sphere_coord_figure.tex} | 
| 672 | %%CNHend | 
| 673 |  | 
| 674 | \subsubsection{Shallow atmosphere approximation} | 
| 675 |  | 
| 676 | Most models are based on the `hydrostatic primitive equations' (HPE's) in | 
| 677 | which the vertical momentum equation is reduced to a statement of | 
| 678 | hydrostatic balance and the `traditional approximation' is made in which the | 
| 679 | Coriolis force is treated approximately and the shallow atmosphere | 
| 680 | approximation is made.\ The MITgcm need not make the `traditional | 
| 681 | approximation'. To be able to support consistent non-hydrostatic forms the | 
| 682 | shallow atmosphere approximation can be relaxed - when dividing through by $ | 
| 683 | r $ in, for example, (\ref{eq:gu-speherical}), we do not replace $r$ by $a$, | 
| 684 | the radius of the earth. | 
| 685 |  | 
| 686 | \subsubsection{Hydrostatic and quasi-hydrostatic forms} | 
| 687 |  | 
| 688 | These are discussed at length in Marshall et al (1997a). | 
| 689 |  | 
| 690 | In the `hydrostatic primitive equations' (\textbf{HPE)} all the underlined | 
| 691 | terms in Eqs. (\ref{eq:gu-speherical} $\rightarrow $\ \ref{eq:gw-spherical}) | 
| 692 | are neglected and `${r}$' is replaced by `$a$', the mean radius of the | 
| 693 | earth. Once the pressure is found at one level - e.g. by inverting a 2-d | 
| 694 | Elliptic equation for $\phi _{s}$ at $r=R_{moving}$ - the pressure can be | 
| 695 | computed at all other levels by integration of the hydrostatic relation, eq( | 
| 696 | \ref{eq:hydrostatic}). | 
| 697 |  | 
| 698 | In the `quasi-hydrostatic' equations (\textbf{QH)} strict balance between | 
| 699 | gravity and vertical pressure gradients is not imposed. The $2\Omega u\cos | 
| 700 | \phi $ Coriolis term are not neglected and are balanced by a non-hydrostatic | 
| 701 | contribution to the pressure field: only the terms underlined twice in Eqs. ( | 
| 702 | \ref{eq:gu-speherical}$\rightarrow $\ \ref{eq:gw-spherical}) are set to zero | 
| 703 | and, simultaneously, the shallow atmosphere approximation is relaxed. In | 
| 704 | \textbf{QH}\ \textit{all} the metric terms are retained and the full | 
| 705 | variation of the radial position of a particle monitored. The \textbf{QH}\ | 
| 706 | vertical momentum equation (\ref{eq:mom-w}) becomes: | 
| 707 |  | 
| 708 | \begin{equation*} | 
| 709 | \frac{\partial \phi _{nh}}{\partial r}=2\Omega u\cos lat | 
| 710 | \end{equation*} | 
| 711 | making a small correction to the hydrostatic pressure. | 
| 712 |  | 
| 713 | \textbf{QH} has good energetic credentials - they are the same as for | 
| 714 | \textbf{HPE}. Importantly, however, it has the same angular momentum | 
| 715 | principle as the full non-hydrostatic model (\textbf{NH)} - see Marshall | 
| 716 | et.al., 1997a. As in \textbf{HPE }only a 2-d elliptic problem need be solved. | 
| 717 |  | 
| 718 | \subsubsection{Non-hydrostatic and quasi-nonhydrostatic forms} | 
| 719 |  | 
| 720 | The MIT model presently supports a full non-hydrostatic ocean isomorph, but | 
| 721 | only a quasi-non-hydrostatic atmospheric isomorph. | 
| 722 |  | 
| 723 | \paragraph{Non-hydrostatic Ocean} | 
| 724 |  | 
| 725 | In the non-hydrostatic ocean model all terms in equations Eqs.(\ref | 
| 726 | {eq:gu-speherical} $\rightarrow $\ \ref{eq:gw-spherical}) are retained. A | 
| 727 | three dimensional elliptic equation must be solved subject to Neumann | 
| 728 | boundary conditions (see below). It is important to note that use of the | 
| 729 | full \textbf{NH} does not admit any new `fast' waves in to the system - the | 
| 730 | incompressible condition eq(\ref{eq:continuous})c has already filtered out | 
| 731 | acoustic modes. It does, however, ensure that the gravity waves are treated | 
| 732 | accurately with an exact dispersion relation. The \textbf{NH} set has a | 
| 733 | complete angular momentum principle and consistent energetics - see White | 
| 734 | and Bromley, 1995; Marshall et.al.\ 1997a. | 
| 735 |  | 
| 736 | \paragraph{Quasi-nonhydrostatic Atmosphere} | 
| 737 |  | 
| 738 | In the non-hydrostatic version of our atmospheric model we approximate $\dot{ | 
| 739 | r}$ in the vertical momentum eqs(\ref{eq:mom-w}) and (\ref{eq:gv-spherical}) | 
| 740 | (but only here) by: | 
| 741 |  | 
| 742 | \begin{equation} | 
| 743 | \dot{r}=\frac{Dp}{Dt}=\frac{1}{g}\frac{D\phi }{Dt}  \label{eq:quasi-nh-w} | 
| 744 | \end{equation} | 
| 745 | where $p_{hy}$ is the hydrostatic pressure. | 
| 746 |  | 
| 747 | \subsubsection{Summary of equation sets supported by model} | 
| 748 |  | 
| 749 | \paragraph{Atmosphere} | 
| 750 |  | 
| 751 | Hydrostatic, and quasi-hydrostatic and quasi non-hydrostatic forms of the | 
| 752 | compressible non-Boussinesq equations in $p-$coordinates are supported. | 
| 753 |  | 
| 754 | \subparagraph{Hydrostatic and quasi-hydrostatic} | 
| 755 |  | 
| 756 | The hydrostatic set is written out in $p-$coordinates in appendix Atmosphere | 
| 757 | - see eq(\ref{eq:atmos-prime}). | 
| 758 |  | 
| 759 | \subparagraph{Quasi-nonhydrostatic} | 
| 760 |  | 
| 761 | A quasi-nonhydrostatic form is also supported. | 
| 762 |  | 
| 763 | \paragraph{Ocean} | 
| 764 |  | 
| 765 | \subparagraph{Hydrostatic and quasi-hydrostatic} | 
| 766 |  | 
| 767 | Hydrostatic, and quasi-hydrostatic forms of the incompressible Boussinesq | 
| 768 | equations in $z-$coordinates are supported. | 
| 769 |  | 
| 770 | \subparagraph{Non-hydrostatic} | 
| 771 |  | 
| 772 | Non-hydrostatic forms of the incompressible Boussinesq equations in $z-$ | 
| 773 | coordinates are supported - see eqs(\ref{eq:ocean-mom}) to (\ref | 
| 774 | {eq:ocean-salt}). | 
| 775 |  | 
| 776 | \subsection{Solution strategy} | 
| 777 |  | 
| 778 | The method of solution employed in the \textbf{HPE}, \textbf{QH} and \textbf{ | 
| 779 | NH} models is summarized in Fig.7. | 
| 780 | \marginpar{ | 
| 781 | Fig.7 Solution strategy} Under all dynamics, a 2-d elliptic equation is | 
| 782 | first solved to find the surface pressure and the hydrostatic pressure at | 
| 783 | any level computed from the weight of fluid above. Under \textbf{HPE} and | 
| 784 | \textbf{QH} dynamics, the horizontal momentum equations are then stepped | 
| 785 | forward and $\dot{r}$ found from continuity. Under \textbf{NH} dynamics a | 
| 786 | 3-d elliptic equation must be solved for the non-hydrostatic pressure before | 
| 787 | stepping forward the horizontal momentum equations; $\dot{r}$ is found by | 
| 788 | stepping forward the vertical momentum equation. | 
| 789 |  | 
| 790 | %%CNHbegin | 
| 791 | %notci%\input{part1/solution_strategy_figure.tex} | 
| 792 | %%CNHend | 
| 793 |  | 
| 794 | There is no penalty in implementing \textbf{QH} over \textbf{HPE} except, of | 
| 795 | course, some complication that goes with the inclusion of $\cos \phi \ $ | 
| 796 | Coriolis terms and the relaxation of the shallow atmosphere approximation. | 
| 797 | But this leads to negligible increase in computation. In \textbf{NH}, in | 
| 798 | contrast, one additional elliptic equation - a three-dimensional one - must | 
| 799 | be inverted for $p_{nh}$. However the `overhead' of the \textbf{NH} model is | 
| 800 | essentially negligible in the hydrostatic limit (see detailed discussion in | 
| 801 | Marshall et al, 1997) resulting in a non-hydrostatic algorithm that, in the | 
| 802 | hydrostatic limit, is as computationally economic as the \textbf{HPEs}. | 
| 803 |  | 
| 804 | \subsection{Finding the pressure field} | 
| 805 |  | 
| 806 | Unlike the prognostic variables $u$, $v$, $w$, $\theta $ and $S$, the | 
| 807 | pressure field must be obtained diagnostically. We proceed, as before, by | 
| 808 | dividing the total (pressure/geo) potential in to three parts, a surface | 
| 809 | part, $\phi _{s}(x,y)$, a hydrostatic part $\phi _{hyd}(x,y,r)$ and a | 
| 810 | non-hydrostatic part $\phi _{nh}(x,y,r)$, as in (\ref{eq:phi-split}), and | 
| 811 | writing the momentum equation as in (\ref{eq:mom-h}). | 
| 812 |  | 
| 813 | \subsubsection{Hydrostatic pressure} | 
| 814 |  | 
| 815 | Hydrostatic pressure is obtained by integrating (\ref{eq:hydrostatic}) | 
| 816 | vertically from $r=R_{o}$ where $\phi _{hyd}(r=R_{o})=0$, to yield: | 
| 817 |  | 
| 818 | \begin{equation*} | 
| 819 | \int_{r}^{R_{o}}\frac{\partial \phi _{hyd}}{\partial r}dr=\left[ \phi _{hyd} | 
| 820 | \right] _{r}^{R_{o}}=\int_{r}^{R_{o}}-bdr | 
| 821 | \end{equation*} | 
| 822 | and so | 
| 823 |  | 
| 824 | \begin{equation} | 
| 825 | \phi _{hyd}(x,y,r)=\int_{r}^{R_{o}}bdr  \label{eq:hydro-phi} | 
| 826 | \end{equation} | 
| 827 |  | 
| 828 | The model can be easily modified to accommodate a loading term (e.g | 
| 829 | atmospheric pressure pushing down on the ocean's surface) by setting: | 
| 830 |  | 
| 831 | \begin{equation} | 
| 832 | \phi _{hyd}(r=R_{o})=loading  \label{eq:loading} | 
| 833 | \end{equation} | 
| 834 |  | 
| 835 | \subsubsection{Surface pressure} | 
| 836 |  | 
| 837 | The surface pressure equation can be obtained by integrating continuity, ( | 
| 838 | \ref{eq:continuous})c, vertically from $r=R_{fixed}$ to $r=R_{moving}$ | 
| 839 |  | 
| 840 | \begin{equation*} | 
| 841 | \int_{R_{fixed}}^{R_{moving}}\left( \mathbf{\nabla }_{h}\cdot \vec{\mathbf{v} | 
| 842 | }_{h}+\partial _{r}\dot{r}\right) dr=0 | 
| 843 | \end{equation*} | 
| 844 |  | 
| 845 | Thus: | 
| 846 |  | 
| 847 | \begin{equation*} | 
| 848 | \frac{\partial \eta }{\partial t}+\vec{\mathbf{v}}.\nabla \eta | 
| 849 | +\int_{R_{fixed}}^{R_{moving}}\mathbf{\nabla }_{h}\cdot \vec{\mathbf{v}} | 
| 850 | _{h}dr=0 | 
| 851 | \end{equation*} | 
| 852 | where $\eta =R_{moving}-R_{o}$ is the free-surface $r$-anomaly in units of $ | 
| 853 | r $. The above can be rearranged to yield, using Leibnitz's theorem: | 
| 854 |  | 
| 855 | \begin{equation} | 
| 856 | \frac{\partial \eta }{\partial t}+\mathbf{\nabla }_{h}\cdot | 
| 857 | \int_{R_{fixed}}^{R_{moving}}\vec{\mathbf{v}}_{h}dr=\text{source} | 
| 858 | \label{eq:free-surface} | 
| 859 | \end{equation} | 
| 860 | where we have incorporated a source term. | 
| 861 |  | 
| 862 | Whether $\phi $ is pressure (ocean model, $p/\rho _{c}$) or geopotential | 
| 863 | (atmospheric model), in (\ref{mtm-split}), the horizontal gradient term can | 
| 864 | be written | 
| 865 | \begin{equation} | 
| 866 | \mathbf{\nabla }_{h}\phi _{s}=\mathbf{\nabla }_{h}\left( b_{s}\eta \right) | 
| 867 | \label{eq:phi-surf} | 
| 868 | \end{equation} | 
| 869 | where $b_{s}$ is the buoyancy at the surface. | 
| 870 |  | 
| 871 | In the hydrostatic limit ($\epsilon _{nh}=0$), Eqs(\ref{eq:mom-h}), (\ref | 
| 872 | {eq:free-surface}) and (\ref{eq:phi-surf}) can be solved by inverting a 2-d | 
| 873 | elliptic equation for $\phi _{s}$ as described in Chapter 2. Both `free | 
| 874 | surface' and `rigid lid' approaches are available. | 
| 875 |  | 
| 876 | \subsubsection{Non-hydrostatic pressure} | 
| 877 |  | 
| 878 | Taking the horizontal divergence of (\ref{hor-mtm}) and adding $\frac{ | 
| 879 | \partial }{\partial r}$ of (\ref{vertmtm}), invoking the continuity equation | 
| 880 | (\ref{incompressible}), we deduce that: | 
| 881 |  | 
| 882 | \begin{equation} | 
| 883 | \nabla _{3}^{2}\phi _{nh}=\nabla .\vec{\mathbf{G}}_{\vec{v}}-\left( \mathbf{ | 
| 884 | \nabla }_{h}^{2}\phi _{s}+\mathbf{\nabla }^{2}\phi _{hyd}\right) =\nabla . | 
| 885 | \vec{\mathbf{F}}  \label{eq:3d-invert} | 
| 886 | \end{equation} | 
| 887 |  | 
| 888 | For a given rhs this 3-d elliptic equation must be inverted for $\phi _{nh}$ | 
| 889 | subject to appropriate choice of boundary conditions. This method is usually | 
| 890 | called \textit{The Pressure Method} [Harlow and Welch, 1965; Williams, 1969; | 
| 891 | Potter, 1976]. In the hydrostatic primitive equations case (\textbf{HPE}), | 
| 892 | the 3-d problem does not need to be solved. | 
| 893 |  | 
| 894 | \paragraph{Boundary Conditions} | 
| 895 |  | 
| 896 | We apply the condition of no normal flow through all solid boundaries - the | 
| 897 | coasts (in the ocean) and the bottom: | 
| 898 |  | 
| 899 | \begin{equation} | 
| 900 | \vec{\mathbf{v}}.\widehat{n}=0  \label{nonormalflow} | 
| 901 | \end{equation} | 
| 902 | where $\widehat{n}$ is a vector of unit length normal to the boundary. The | 
| 903 | kinematic condition (\ref{nonormalflow}) is also applied to the vertical | 
| 904 | velocity at $r=R_{moving}$. No-slip $\left( v_{T}=0\right) \ $or slip $ | 
| 905 | \left( \partial v_{T}/\partial n=0\right) \ $conditions are employed on the | 
| 906 | tangential component of velocity, $v_{T}$, at all solid boundaries, | 
| 907 | depending on the form chosen for the dissipative terms in the momentum | 
| 908 | equations - see below. | 
| 909 |  | 
| 910 | Eq.(\ref{nonormalflow}) implies, making use of (\ref{mtm-split}), that: | 
| 911 |  | 
| 912 | \begin{equation} | 
| 913 | \widehat{n}.\nabla \phi _{nh}=\widehat{n}.\vec{\mathbf{F}} | 
| 914 | \label{eq:inhom-neumann-nh} | 
| 915 | \end{equation} | 
| 916 | where | 
| 917 |  | 
| 918 | \begin{equation*} | 
| 919 | \vec{\mathbf{F}}=\vec{\mathbf{G}}_{\vec{v}}-\left( \mathbf{\nabla }_{h}\phi | 
| 920 | _{s}+\mathbf{\nabla }\phi _{hyd}\right) | 
| 921 | \end{equation*} | 
| 922 | presenting inhomogeneous Neumann boundary conditions to the Elliptic problem | 
| 923 | (\ref{eq:3d-invert}). As shown, for example, by Williams (1969), one can | 
| 924 | exploit classical 3D potential theory and, by introducing an appropriately | 
| 925 | chosen $\delta $-function sheet of `source-charge', replace the | 
| 926 | inhomogeneous boundary condition on pressure by a homogeneous one. The | 
| 927 | source term $rhs$ in (\ref{eq:3d-invert}) is the divergence of the vector $ | 
| 928 | \vec{\mathbf{F}}.$ By simultaneously setting $ | 
| 929 | \begin{array}{l} | 
| 930 | \widehat{n}.\vec{\mathbf{F}} | 
| 931 | \end{array} | 
| 932 | =0$\ and $\widehat{n}.\nabla \phi _{nh}=0\ $on the boundary the following | 
| 933 | self-consistent but simpler homogenized Elliptic problem is obtained: | 
| 934 |  | 
| 935 | \begin{equation*} | 
| 936 | \nabla ^{2}\phi _{nh}=\nabla .\widetilde{\vec{\mathbf{F}}}\qquad | 
| 937 | \end{equation*} | 
| 938 | where $\widetilde{\vec{\mathbf{F}}}$ is a modified $\vec{\mathbf{F}}$ such | 
| 939 | that $\widetilde{\vec{\mathbf{F}}}.\widehat{n}=0$. As is implied by (\ref | 
| 940 | {eq:inhom-neumann-nh}) the modified boundary condition becomes: | 
| 941 |  | 
| 942 | \begin{equation} | 
| 943 | \widehat{n}.\nabla \phi _{nh}=0  \label{eq:hom-neumann-nh} | 
| 944 | \end{equation} | 
| 945 |  | 
| 946 | If the flow is `close' to hydrostatic balance then the 3-d inversion | 
| 947 | converges rapidly because $\phi _{nh}\ $is then only a small correction to | 
| 948 | the hydrostatic pressure field (see the discussion in Marshall et al, a,b). | 
| 949 |  | 
| 950 | The solution $\phi _{nh}\ $to (\ref{eq:3d-invert}) and (\ref{homneuman}) | 
| 951 | does not vanish at $r=R_{moving}$, and so refines the pressure there. | 
| 952 |  | 
| 953 | \subsection{Forcing/dissipation} | 
| 954 |  | 
| 955 | \subsubsection{Forcing} | 
| 956 |  | 
| 957 | The forcing terms $\mathcal{F}$ on the rhs of the equations are provided by | 
| 958 | `physics packages' described in detail in chapter ??. | 
| 959 |  | 
| 960 | \subsubsection{Dissipation} | 
| 961 |  | 
| 962 | \paragraph{Momentum} | 
| 963 |  | 
| 964 | Many forms of momentum dissipation are available in the model. Laplacian and | 
| 965 | biharmonic frictions are commonly used: | 
| 966 |  | 
| 967 | \begin{equation} | 
| 968 | D_{V}=A_{h}\nabla _{h}^{2}v+A_{v}\frac{\partial ^{2}v}{\partial z^{2}} | 
| 969 | +A_{4}\nabla _{h}^{4}v  \label{eq:dissipation} | 
| 970 | \end{equation} | 
| 971 | where $A_{h}$ and $A_{v}\ $are (constant) horizontal and vertical viscosity | 
| 972 | coefficients and $A_{4}\ $is the horizontal coefficient for biharmonic | 
| 973 | friction. These coefficients are the same for all velocity components. | 
| 974 |  | 
| 975 | \paragraph{Tracers} | 
| 976 |  | 
| 977 | The mixing terms for the temperature and salinity equations have a similar | 
| 978 | form to that of momentum except that the diffusion tensor can be | 
| 979 | non-diagonal and have varying coefficients. $\qquad $ | 
| 980 | \begin{equation} | 
| 981 | D_{T,S}=\nabla .[\underline{\underline{K}}\nabla (T,S)]+K_{4}\nabla | 
| 982 | _{h}^{4}(T,S)  \label{eq:diffusion} | 
| 983 | \end{equation} | 
| 984 | where $\underline{\underline{K}}\ $is the diffusion tensor and the $K_{4}\ $ | 
| 985 | horizontal coefficient for biharmonic diffusion. In the simplest case where | 
| 986 | the subgrid-scale fluxes of heat and salt are parameterized with constant | 
| 987 | horizontal and vertical diffusion coefficients, $\underline{\underline{K}}$, | 
| 988 | reduces to a diagonal matrix with constant coefficients: | 
| 989 |  | 
| 990 | \begin{equation} | 
| 991 | \qquad \qquad \qquad \qquad K=\left( | 
| 992 | \begin{array}{ccc} | 
| 993 | K_{h} & 0 & 0 \\ | 
| 994 | 0 & K_{h} & 0 \\ | 
| 995 | 0 & 0 & K_{v} | 
| 996 | \end{array} | 
| 997 | \right) \qquad \qquad \qquad  \label{eq:diagonal-diffusion-tensor} | 
| 998 | \end{equation} | 
| 999 | where $K_{h}\ $and $K_{v}\ $are the horizontal and vertical diffusion | 
| 1000 | coefficients. These coefficients are the same for all tracers (temperature, | 
| 1001 | salinity ... ). | 
| 1002 |  | 
| 1003 | \subsection{Vector invariant form} | 
| 1004 |  | 
| 1005 | For some purposes it is advantageous to write momentum advection in eq(\ref | 
| 1006 | {hor-mtm}) and (\ref{vertmtm}) in the (so-called) `vector invariant' form: | 
| 1007 |  | 
| 1008 | \begin{equation} | 
| 1009 | \frac{D\vec{\mathbf{v}}}{Dt}=\frac{\partial \vec{\mathbf{v}}}{\partial t} | 
| 1010 | +\left( \nabla \times \vec{\mathbf{v}}\right) \times \vec{\mathbf{v}}+\nabla | 
| 1011 | \left[ \frac{1}{2}(\vec{\mathbf{v}}\cdot \vec{\mathbf{v}})\right] | 
| 1012 | \label{eq:vi-identity} | 
| 1013 | \end{equation} | 
| 1014 | This permits alternative numerical treatments of the non-linear terms based | 
| 1015 | on their representation as a vorticity flux. Because gradients of coordinate | 
| 1016 | vectors no longer appear on the rhs of (\ref{eq:vi-identity}), explicit | 
| 1017 | representation of the metric terms in (\ref{eq:gu-speherical}), (\ref | 
| 1018 | {eq:gv-spherical}) and (\ref{eq:gw-spherical}), can be avoided: information | 
| 1019 | about the geometry is contained in the areas and lengths of the volumes used | 
| 1020 | to discretize the model. | 
| 1021 |  | 
| 1022 | \subsection{Adjoint} | 
| 1023 |  | 
| 1024 | Tangent linear and adjoint counterparts of the forward model and described | 
| 1025 | in Chapter 5. | 
| 1026 |  | 
| 1027 | % $Header: /u/gcmpack/mitgcmdoc/part1/manual.tex,v 1.3 2001/10/10 16:48:45 cnh Exp $ | 
| 1028 | % $Name:  $ | 
| 1029 |  | 
| 1030 | \section{Appendix ATMOSPHERE} | 
| 1031 |  | 
| 1032 | \subsection{Hydrostatic Primitive Equations for the Atmosphere in pressure | 
| 1033 | coordinates} | 
| 1034 |  | 
| 1035 | \label{sect-hpe-p} | 
| 1036 |  | 
| 1037 | The hydrostatic primitive equations (HPEs) in p-coordinates are: | 
| 1038 | \begin{eqnarray} | 
| 1039 | \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}} | 
| 1040 | _{h}+\mathbf{\nabla }_{p}\phi &=&\vec{\mathbf{\mathcal{F}}} | 
| 1041 | \label{eq:atmos-mom} \\ | 
| 1042 | \frac{\partial \phi }{\partial p}+\alpha &=&0  \label{eq-p-hydro-start} \\ | 
| 1043 | \mathbf{\nabla }_{p}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \omega }{ | 
| 1044 | \partial p} &=&0  \label{eq:atmos-cont} \\ | 
| 1045 | p\alpha &=&RT  \label{eq:atmos-eos} \\ | 
| 1046 | c_{v}\frac{DT}{Dt}+p\frac{D\alpha }{Dt} &=&\mathcal{Q}  \label{eq:atmos-heat} | 
| 1047 | \end{eqnarray} | 
| 1048 | where $\vec{\mathbf{v}}_{h}=(u,v,0)$ is the `horizontal' (on pressure | 
| 1049 | surfaces) component of velocity,$\frac{D}{Dt}=\vec{\mathbf{v}}_{h}\cdot | 
| 1050 | \mathbf{\nabla }_{p}+\omega \frac{\partial }{\partial p}$ is the total | 
| 1051 | derivative, $f=2\Omega \sin lat$ is the Coriolis parameter, $\phi =gz$ is | 
| 1052 | the geopotential, $\alpha =1/\rho $ is the specific volume, $\omega =\frac{Dp | 
| 1053 | }{Dt}$ is the vertical velocity in the $p-$coordinate. Equation(\ref | 
| 1054 | {eq:atmos-heat}) is the first law of thermodynamics where internal energy $ | 
| 1055 | e=c_{v}T$, $T$ is temperature, $Q$ is the rate of heating per unit mass and $ | 
| 1056 | p\frac{D\alpha }{Dt}$ is the work done by the fluid in compressing. | 
| 1057 |  | 
| 1058 | It is convenient to cast the heat equation in terms of potential temperature | 
| 1059 | $\theta $ so that it looks more like a generic conservation law. | 
| 1060 | Differentiating (\ref{eq:atmos-eos}) we get: | 
| 1061 | \begin{equation*} | 
| 1062 | p\frac{D\alpha }{Dt}+\alpha \frac{Dp}{Dt}=R\frac{DT}{Dt} | 
| 1063 | \end{equation*} | 
| 1064 | which, when added to the heat equation (\ref{eq:atmos-heat}) and using $ | 
| 1065 | c_{p}=c_{v}+R$, gives: | 
| 1066 | \begin{equation} | 
| 1067 | c_{p}\frac{DT}{Dt}-\alpha \frac{Dp}{Dt}=\mathcal{Q} | 
| 1068 | \label{eq-p-heat-interim} | 
| 1069 | \end{equation} | 
| 1070 | Potential temperature is defined: | 
| 1071 | \begin{equation} | 
| 1072 | \theta =T(\frac{p_{c}}{p})^{\kappa }  \label{eq:potential-temp} | 
| 1073 | \end{equation} | 
| 1074 | where $p_{c}$ is a reference pressure and $\kappa =R/c_{p}$. For convenience | 
| 1075 | we will make use of the Exner function $\Pi (p)$ which defined by: | 
| 1076 | \begin{equation} | 
| 1077 | \Pi (p)=c_{p}(\frac{p}{p_{c}})^{\kappa }  \label{Exner} | 
| 1078 | \end{equation} | 
| 1079 | The following relations will be useful and are easily expressed in terms of | 
| 1080 | the Exner function: | 
| 1081 | \begin{equation*} | 
| 1082 | c_{p}T=\Pi \theta \;\;;\;\;\frac{\partial \Pi }{\partial p}=\frac{\kappa \Pi | 
| 1083 | }{p}\;\;;\;\;\alpha =\frac{\kappa \Pi \theta }{p}=\frac{\partial \ \Pi }{ | 
| 1084 | \partial p}\theta \;\;;\;\;\frac{D\Pi }{Dt}=\frac{\partial \Pi }{\partial p} | 
| 1085 | \frac{Dp}{Dt} | 
| 1086 | \end{equation*} | 
| 1087 | where $b=\frac{\partial \ \Pi }{\partial p}\theta $ is the buoyancy. | 
| 1088 |  | 
| 1089 | The heat equation is obtained by noting that | 
| 1090 | \begin{equation*} | 
| 1091 | c_{p}\frac{DT}{Dt}=\frac{D(\Pi \theta )}{Dt}=\Pi \frac{D\theta }{Dt}+\theta | 
| 1092 | \frac{D\Pi }{Dt}=\Pi \frac{D\theta }{Dt}+\alpha \frac{Dp}{Dt} | 
| 1093 | \end{equation*} | 
| 1094 | and on substituting into (\ref{eq-p-heat-interim}) gives: | 
| 1095 | \begin{equation} | 
| 1096 | \Pi \frac{D\theta }{Dt}=\mathcal{Q} | 
| 1097 | \label{eq:potential-temperature-equation} | 
| 1098 | \end{equation} | 
| 1099 | which is in conservative form. | 
| 1100 |  | 
| 1101 | For convenience in the model we prefer to step forward (\ref | 
| 1102 | {eq:potential-temperature-equation}) rather than (\ref{eq:atmos-heat}). | 
| 1103 |  | 
| 1104 | \subsubsection{Boundary conditions} | 
| 1105 |  | 
| 1106 | The upper and lower boundary conditions are : | 
| 1107 | \begin{eqnarray} | 
| 1108 | \mbox{at the top:}\;\;p=0 &&\text{, }\omega =\frac{Dp}{Dt}=0 \\ | 
| 1109 | \mbox{at the surface:}\;\;p=p_{s} &&\text{, }\phi =\phi _{topo}=g~Z_{topo} | 
| 1110 | \label{eq:boundary-condition-atmosphere} | 
| 1111 | \end{eqnarray} | 
| 1112 | In $p$-coordinates, the upper boundary acts like a solid boundary ($\omega | 
| 1113 | =0 $); in $z$-coordinates and the lower boundary is analogous to a free | 
| 1114 | surface ($\phi $ is imposed and $\omega \neq 0$). | 
| 1115 |  | 
| 1116 | \subsubsection{Splitting the geo-potential} | 
| 1117 |  | 
| 1118 | For the purposes of initialization and reducing round-off errors, the model | 
| 1119 | deals with perturbations from reference (or ``standard'') profiles. For | 
| 1120 | example, the hydrostatic geopotential associated with the resting atmosphere | 
| 1121 | is not dynamically relevant and can therefore be subtracted from the | 
| 1122 | equations. The equations written in terms of perturbations are obtained by | 
| 1123 | substituting the following definitions into the previous model equations: | 
| 1124 | \begin{eqnarray} | 
| 1125 | \theta &=&\theta _{o}+\theta ^{\prime }  \label{eq:atmos-ref-prof-theta} \\ | 
| 1126 | \alpha &=&\alpha _{o}+\alpha ^{\prime }  \label{eq:atmos-ref-prof-alpha} \\ | 
| 1127 | \phi &=&\phi _{o}+\phi ^{\prime }  \label{eq:atmos-ref-prof-phi} | 
| 1128 | \end{eqnarray} | 
| 1129 | The reference state (indicated by subscript ``0'') corresponds to | 
| 1130 | horizontally homogeneous atmosphere at rest ($\theta _{o},\alpha _{o},\phi | 
| 1131 | _{o}$) with surface pressure $p_{o}(x,y)$ that satisfies $\phi | 
| 1132 | _{o}(p_{o})=g~Z_{topo}$, defined: | 
| 1133 | \begin{eqnarray*} | 
| 1134 | \theta _{o}(p) &=&f^{n}(p) \\ | 
| 1135 | \alpha _{o}(p) &=&\Pi _{p}\theta _{o} \\ | 
| 1136 | \phi _{o}(p) &=&\phi _{topo}-\int_{p_{0}}^{p}\alpha _{o}dp | 
| 1137 | \end{eqnarray*} | 
| 1138 | %\begin{eqnarray*} | 
| 1139 | %\phi'_\alpha & = & \int^p_{p_o} (\alpha_o -\alpha) dp \\ | 
| 1140 | %\phi'_s(x,y,t) & = & \int_{p_o}^{p_s} \alpha dp | 
| 1141 | %\end{eqnarray*} | 
| 1142 |  | 
| 1143 | The final form of the HPE's in p coordinates is then: | 
| 1144 | \begin{eqnarray} | 
| 1145 | \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}} | 
| 1146 | _{h}+\mathbf{\nabla }_{p}\phi ^{\prime } &=&\vec{\mathbf{\mathcal{F}}} \\ | 
| 1147 | \frac{\partial \phi ^{\prime }}{\partial p}+\alpha ^{\prime } &=&0 \\ | 
| 1148 | \mathbf{\nabla }_{p}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial \omega }{ | 
| 1149 | \partial p} &=&0 \\ | 
| 1150 | \frac{\partial \Pi }{\partial p}\theta ^{\prime } &=&\alpha ^{\prime } \\ | 
| 1151 | \frac{D\theta }{Dt} &=&\frac{\mathcal{Q}}{\Pi }  \label{eq:atmos-prime} | 
| 1152 | \end{eqnarray} | 
| 1153 |  | 
| 1154 | % $Header: /u/gcmpack/mitgcmdoc/part1/manual.tex,v 1.3 2001/10/10 16:48:45 cnh Exp $ | 
| 1155 | % $Name:  $ | 
| 1156 |  | 
| 1157 | \section{Appendix OCEAN} | 
| 1158 |  | 
| 1159 | \subsection{Equations of motion for the ocean} | 
| 1160 |  | 
| 1161 | We review here the method by which the standard (Boussinesq, incompressible) | 
| 1162 | HPE's for the ocean written in z-coordinates are obtained. The | 
| 1163 | non-Boussinesq equations for oceanic motion are: | 
| 1164 | \begin{eqnarray} | 
| 1165 | \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}} | 
| 1166 | _{h}+\frac{1}{\rho }\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}} \\ | 
| 1167 | \epsilon _{nh}\frac{Dw}{Dt}+g+\frac{1}{\rho }\frac{\partial p}{\partial z} | 
| 1168 | &=&\epsilon _{nh}\mathcal{F}_{w} \\ | 
| 1169 | \frac{1}{\rho }\frac{D\rho }{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}} | 
| 1170 | _{h}+\frac{\partial w}{\partial z} &=&0 \\ | 
| 1171 | \rho &=&\rho (\theta ,S,p) \\ | 
| 1172 | \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta } \\ | 
| 1173 | \frac{DS}{Dt} &=&\mathcal{Q}_{s}  \label{eq:non-boussinesq} | 
| 1174 | \end{eqnarray} | 
| 1175 | These equations permit acoustics modes, inertia-gravity waves, | 
| 1176 | non-hydrostatic motions, a geostrophic (Rossby) mode and a thermo-haline | 
| 1177 | mode. As written, they cannot be integrated forward consistently - if we | 
| 1178 | step $\rho $ forward in (\ref{eq-zns-cont}), the answer will not be | 
| 1179 | consistent with that obtained by stepping (\ref{eq-zns-heat}) and (\ref | 
| 1180 | {eq-zns-salt}) and then using (\ref{eq-zns-eos}) to yield $\rho $. It is | 
| 1181 | therefore necessary to manipulate the system as follows. Differentiating the | 
| 1182 | EOS (equation of state) gives: | 
| 1183 |  | 
| 1184 | \begin{equation} | 
| 1185 | \frac{D\rho }{Dt}=\left. \frac{\partial \rho }{\partial \theta }\right| | 
| 1186 | _{S,p}\frac{D\theta }{Dt}+\left. \frac{\partial \rho }{\partial S}\right| | 
| 1187 | _{\theta ,p}\frac{DS}{Dt}+\left. \frac{\partial \rho }{\partial p}\right| | 
| 1188 | _{\theta ,S}\frac{Dp}{Dt}  \label{EOSexpansion} | 
| 1189 | \end{equation} | 
| 1190 |  | 
| 1191 | Note that $\frac{\partial \rho }{\partial p}=\frac{1}{c_{s}^{2}}$ is the | 
| 1192 | reciprocal of the sound speed ($c_{s}$) squared. Substituting into \ref | 
| 1193 | {eq-zns-cont} gives: | 
| 1194 | \begin{equation} | 
| 1195 | \frac{1}{\rho c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{ | 
| 1196 | v}}+\partial _{z}w\approx 0  \label{eq-zns-pressure} | 
| 1197 | \end{equation} | 
| 1198 | where we have used an approximation sign to indicate that we have assumed | 
| 1199 | adiabatic motion, dropping the $\frac{D\theta }{Dt}$ and $\frac{DS}{Dt}$. | 
| 1200 | Replacing \ref{eq-zns-cont} with \ref{eq-zns-pressure} yields a system that | 
| 1201 | can be explicitly integrated forward: | 
| 1202 | \begin{eqnarray} | 
| 1203 | \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}} | 
| 1204 | _{h}+\frac{1}{\rho }\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}} | 
| 1205 | \label{eq-cns-hmom} \\ | 
| 1206 | \epsilon _{nh}\frac{Dw}{Dt}+g+\frac{1}{\rho }\frac{\partial p}{\partial z} | 
| 1207 | &=&\epsilon _{nh}\mathcal{F}_{w}  \label{eq-cns-hydro} \\ | 
| 1208 | \frac{1}{\rho c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{\mathbf{ | 
| 1209 | v}}_{h}+\frac{\partial w}{\partial z} &=&0  \label{eq-cns-cont} \\ | 
| 1210 | \rho &=&\rho (\theta ,S,p)  \label{eq-cns-eos} \\ | 
| 1211 | \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta }  \label{eq-cns-heat} \\ | 
| 1212 | \frac{DS}{Dt} &=&\mathcal{Q}_{s}  \label{eq-cns-salt} | 
| 1213 | \end{eqnarray} | 
| 1214 |  | 
| 1215 | \subsubsection{Compressible z-coordinate equations} | 
| 1216 |  | 
| 1217 | Here we linearize the acoustic modes by replacing $\rho $ with $\rho _{o}(z)$ | 
| 1218 | wherever it appears in a product (ie. non-linear term) - this is the | 
| 1219 | `Boussinesq assumption'. The only term that then retains the full variation | 
| 1220 | in $\rho $ is the gravitational acceleration: | 
| 1221 | \begin{eqnarray} | 
| 1222 | \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}} | 
| 1223 | _{h}+\frac{1}{\rho _{o}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}} | 
| 1224 | \label{eq-zcb-hmom} \\ | 
| 1225 | \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{o}}+\frac{1}{\rho _{o}} | 
| 1226 | \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w} | 
| 1227 | \label{eq-zcb-hydro} \\ | 
| 1228 | \frac{1}{\rho _{o}c_{s}^{2}}\frac{Dp}{Dt}+\mathbf{\nabla }_{z}\cdot \vec{ | 
| 1229 | \mathbf{v}}_{h}+\frac{\partial w}{\partial z} &=&0  \label{eq-zcb-cont} \\ | 
| 1230 | \rho &=&\rho (\theta ,S,p)  \label{eq-zcb-eos} \\ | 
| 1231 | \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta }  \label{eq-zcb-heat} \\ | 
| 1232 | \frac{DS}{Dt} &=&\mathcal{Q}_{s}  \label{eq-zcb-salt} | 
| 1233 | \end{eqnarray} | 
| 1234 | These equations still retain acoustic modes. But, because the | 
| 1235 | ``compressible'' terms are linearized, the pressure equation \ref | 
| 1236 | {eq-zcb-cont} can be integrated implicitly with ease (the time-dependent | 
| 1237 | term appears as a Helmholtz term in the non-hydrostatic pressure equation). | 
| 1238 | These are the \emph{truly} compressible Boussinesq equations. Note that the | 
| 1239 | EOS must have the same pressure dependency as the linearized pressure term, | 
| 1240 | ie. $\left. \frac{\partial \rho }{\partial p}\right| _{\theta ,S}=\frac{1}{ | 
| 1241 | c_{s}^{2}}$, for consistency. | 
| 1242 |  | 
| 1243 | \subsubsection{`Anelastic' z-coordinate equations} | 
| 1244 |  | 
| 1245 | The anelastic approximation filters the acoustic mode by removing the | 
| 1246 | time-dependency in the continuity (now pressure-) equation (\ref{eq-zcb-cont} | 
| 1247 | ). This could be done simply by noting that $\frac{Dp}{Dt}\approx -g\rho _{o} | 
| 1248 | \frac{Dz}{Dt}=-g\rho _{o}w$, but this leads to an inconsistency between | 
| 1249 | continuity and EOS. A better solution is to change the dependency on | 
| 1250 | pressure in the EOS by splitting the pressure into a reference function of | 
| 1251 | height and a perturbation: | 
| 1252 | \begin{equation*} | 
| 1253 | \rho =\rho (\theta ,S,p_{o}(z)+\epsilon _{s}p^{\prime }) | 
| 1254 | \end{equation*} | 
| 1255 | Remembering that the term $\frac{Dp}{Dt}$ in continuity comes from | 
| 1256 | differentiating the EOS, the continuity equation then becomes: | 
| 1257 | \begin{equation*} | 
| 1258 | \frac{1}{\rho _{o}c_{s}^{2}}\left( \frac{Dp_{o}}{Dt}+\epsilon _{s}\frac{ | 
| 1259 | Dp^{\prime }}{Dt}\right) +\mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+ | 
| 1260 | \frac{\partial w}{\partial z}=0 | 
| 1261 | \end{equation*} | 
| 1262 | If the time- and space-scales of the motions of interest are longer than | 
| 1263 | those of acoustic modes, then $\frac{Dp^{\prime }}{Dt}<<(\frac{Dp_{o}}{Dt}, | 
| 1264 | \mathbf{\nabla }\cdot \vec{\mathbf{v}}_{h})$ in the continuity equations and | 
| 1265 | $\left. \frac{\partial \rho }{\partial p}\right| _{\theta ,S}\frac{ | 
| 1266 | Dp^{\prime }}{Dt}<<\left. \frac{\partial \rho }{\partial p}\right| _{\theta | 
| 1267 | ,S}\frac{Dp_{o}}{Dt}$ in the EOS (\ref{EOSexpansion}). Thus we set $\epsilon | 
| 1268 | _{s}=0$, removing the dependency on $p^{\prime }$ in the continuity equation | 
| 1269 | and EOS. Expanding $\frac{Dp_{o}(z)}{Dt}=-g\rho _{o}w$ then leads to the | 
| 1270 | anelastic continuity equation: | 
| 1271 | \begin{equation} | 
| 1272 | \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z}- | 
| 1273 | \frac{g}{c_{s}^{2}}w=0  \label{eq-za-cont1} | 
| 1274 | \end{equation} | 
| 1275 | A slightly different route leads to the quasi-Boussinesq continuity equation | 
| 1276 | where we use the scaling $\frac{\partial \rho ^{\prime }}{\partial t}+ | 
| 1277 | \mathbf{\nabla }_{3}\cdot \rho ^{\prime }\vec{\mathbf{v}}<<\mathbf{\nabla } | 
| 1278 | _{3}\cdot \rho _{o}\vec{\mathbf{v}}$ yielding: | 
| 1279 | \begin{equation} | 
| 1280 | \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{1}{\rho _{o}}\frac{ | 
| 1281 | \partial \left( \rho _{o}w\right) }{\partial z}=0  \label{eq-za-cont2} | 
| 1282 | \end{equation} | 
| 1283 | Equations \ref{eq-za-cont1} and \ref{eq-za-cont2} are in fact the same | 
| 1284 | equation if: | 
| 1285 | \begin{equation} | 
| 1286 | \frac{1}{\rho _{o}}\frac{\partial \rho _{o}}{\partial z}=\frac{-g}{c_{s}^{2}} | 
| 1287 | \end{equation} | 
| 1288 | Again, note that if $\rho _{o}$ is evaluated from prescribed $\theta _{o}$ | 
| 1289 | and $S_{o}$ profiles, then the EOS dependency on $p_{o}$ and the term $\frac{ | 
| 1290 | g}{c_{s}^{2}}$ in continuity should be referred to those same profiles. The | 
| 1291 | full set of `quasi-Boussinesq' or `anelastic' equations for the ocean are | 
| 1292 | then: | 
| 1293 | \begin{eqnarray} | 
| 1294 | \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}} | 
| 1295 | _{h}+\frac{1}{\rho _{o}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}} | 
| 1296 | \label{eq-zab-hmom} \\ | 
| 1297 | \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{o}}+\frac{1}{\rho _{o}} | 
| 1298 | \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w} | 
| 1299 | \label{eq-zab-hydro} \\ | 
| 1300 | \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{1}{\rho _{o}}\frac{ | 
| 1301 | \partial \left( \rho _{o}w\right) }{\partial z} &=&0  \label{eq-zab-cont} \\ | 
| 1302 | \rho &=&\rho (\theta ,S,p_{o}(z))  \label{eq-zab-eos} \\ | 
| 1303 | \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta }  \label{eq-zab-heat} \\ | 
| 1304 | \frac{DS}{Dt} &=&\mathcal{Q}_{s}  \label{eq-zab-salt} | 
| 1305 | \end{eqnarray} | 
| 1306 |  | 
| 1307 | \subsubsection{Incompressible z-coordinate equations} | 
| 1308 |  | 
| 1309 | Here, the objective is to drop the depth dependence of $\rho _{o}$ and so, | 
| 1310 | technically, to also remove the dependence of $\rho $ on $p_{o}$. This would | 
| 1311 | yield the ``truly'' incompressible Boussinesq equations: | 
| 1312 | \begin{eqnarray} | 
| 1313 | \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}} | 
| 1314 | _{h}+\frac{1}{\rho _{c}}\mathbf{\nabla }_{z}p &=&\vec{\mathbf{\mathcal{F}}} | 
| 1315 | \label{eq-ztb-hmom} \\ | 
| 1316 | \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho }{\rho _{c}}+\frac{1}{\rho _{c}} | 
| 1317 | \frac{\partial p}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w} | 
| 1318 | \label{eq-ztb-hydro} \\ | 
| 1319 | \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z} | 
| 1320 | &=&0  \label{eq-ztb-cont} \\ | 
| 1321 | \rho &=&\rho (\theta ,S)  \label{eq-ztb-eos} \\ | 
| 1322 | \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta }  \label{eq-ztb-heat} \\ | 
| 1323 | \frac{DS}{Dt} &=&\mathcal{Q}_{s}  \label{eq-ztb-salt} | 
| 1324 | \end{eqnarray} | 
| 1325 | where $\rho _{c}$ is a constant reference density of water. | 
| 1326 |  | 
| 1327 | \subsubsection{Compressible non-divergent equations} | 
| 1328 |  | 
| 1329 | The above ``incompressible'' equations are incompressible in both the flow | 
| 1330 | and the density. In many oceanic applications, however, it is important to | 
| 1331 | retain compressibility effects in the density. To do this we must split the | 
| 1332 | density thus: | 
| 1333 | \begin{equation*} | 
| 1334 | \rho =\rho _{o}+\rho ^{\prime } | 
| 1335 | \end{equation*} | 
| 1336 | We then assert that variations with depth of $\rho _{o}$ are unimportant | 
| 1337 | while the compressible effects in $\rho ^{\prime }$ are: | 
| 1338 | \begin{equation*} | 
| 1339 | \rho _{o}=\rho _{c} | 
| 1340 | \end{equation*} | 
| 1341 | \begin{equation*} | 
| 1342 | \rho ^{\prime }=\rho (\theta ,S,p_{o}(z))-\rho _{o} | 
| 1343 | \end{equation*} | 
| 1344 | This then yields what we can call the semi-compressible Boussinesq | 
| 1345 | equations: | 
| 1346 | \begin{eqnarray} | 
| 1347 | \frac{D\vec{\mathbf{v}}_{h}}{Dt}+f\hat{\mathbf{k}}\times \vec{\mathbf{v}} | 
| 1348 | _{h}+\frac{1}{\rho _{c}}\mathbf{\nabla }_{z}p^{\prime } &=&\vec{\mathbf{ | 
| 1349 | \mathcal{F}}}  \label{eq:ocean-mom} \\ | 
| 1350 | \epsilon _{nh}\frac{Dw}{Dt}+\frac{g\rho ^{\prime }}{\rho _{c}}+\frac{1}{\rho | 
| 1351 | _{c}}\frac{\partial p^{\prime }}{\partial z} &=&\epsilon _{nh}\mathcal{F}_{w} | 
| 1352 | \label{eq:ocean-wmom} \\ | 
| 1353 | \mathbf{\nabla }_{z}\cdot \vec{\mathbf{v}}_{h}+\frac{\partial w}{\partial z} | 
| 1354 | &=&0  \label{eq:ocean-cont} \\ | 
| 1355 | \rho ^{\prime } &=&\rho (\theta ,S,p_{o}(z))-\rho _{c}  \label{eq:ocean-eos} | 
| 1356 | \\ | 
| 1357 | \frac{D\theta }{Dt} &=&\mathcal{Q}_{\theta }  \label{eq:ocean-theta} \\ | 
| 1358 | \frac{DS}{Dt} &=&\mathcal{Q}_{s}  \label{eq:ocean-salt} | 
| 1359 | \end{eqnarray} | 
| 1360 | Note that the hydrostatic pressure of the resting fluid, including that | 
| 1361 | associated with $\rho _{c}$, is subtracted out since it has no effect on the | 
| 1362 | dynamics. | 
| 1363 |  | 
| 1364 | Though necessary, the assumptions that go into these equations are messy | 
| 1365 | since we essentially assume a different EOS for the reference density and | 
| 1366 | the perturbation density. Nevertheless, it is the hydrostatic ($\epsilon | 
| 1367 | _{nh}=0$ form of these equations that are used throughout the ocean modeling | 
| 1368 | community and referred to as the primitive equations (HPE). | 
| 1369 |  | 
| 1370 | % $Header: /u/gcmpack/mitgcmdoc/part1/manual.tex,v 1.3 2001/10/10 16:48:45 cnh Exp $ | 
| 1371 | % $Name:  $ | 
| 1372 |  | 
| 1373 | \section{Appendix:OPERATORS} | 
| 1374 |  | 
| 1375 | \subsection{Coordinate systems} | 
| 1376 |  | 
| 1377 | \subsubsection{Spherical coordinates} | 
| 1378 |  | 
| 1379 | In spherical coordinates, the velocity components in the zonal, meridional | 
| 1380 | and vertical direction respectively, are given by (see Fig.2) : | 
| 1381 |  | 
| 1382 | \begin{equation*} | 
| 1383 | u=r\cos \phi \frac{D\lambda }{Dt} | 
| 1384 | \end{equation*} | 
| 1385 |  | 
| 1386 | \begin{equation*} | 
| 1387 | v=r\frac{D\phi }{Dt}\qquad | 
| 1388 | \end{equation*} | 
| 1389 | $\qquad \qquad \qquad \qquad $ | 
| 1390 |  | 
| 1391 | \begin{equation*} | 
| 1392 | \dot{r}=\frac{Dr}{Dt} | 
| 1393 | \end{equation*} | 
| 1394 |  | 
| 1395 | Here $\phi $ is the latitude, $\lambda $ the longitude, $r$ the radial | 
| 1396 | distance of the particle from the center of the earth, $\Omega $ is the | 
| 1397 | angular speed of rotation of the Earth and $D/Dt$ is the total derivative. | 
| 1398 |  | 
| 1399 | The `grad' ($\nabla $) and `div' ($\nabla $.) operators are defined by, in | 
| 1400 | spherical coordinates: | 
| 1401 |  | 
| 1402 | \begin{equation*} | 
| 1403 | \nabla \equiv \left( \frac{1}{r\cos \phi }\frac{\partial }{\partial \lambda } | 
| 1404 | ,\frac{1}{r}\frac{\partial }{\partial \phi },\frac{\partial }{\partial r} | 
| 1405 | \right) | 
| 1406 | \end{equation*} | 
| 1407 |  | 
| 1408 | \begin{equation*} | 
| 1409 | \nabla .v\equiv \frac{1}{r\cos \phi }\left\{ \frac{\partial u}{\partial | 
| 1410 | \lambda }+\frac{\partial }{\partial \phi }\left( v\cos \phi \right) \right\} | 
| 1411 | +\frac{1}{r^{2}}\frac{\partial \left( r^{2}\dot{r}\right) }{\partial r} | 
| 1412 | \end{equation*} | 
| 1413 |  | 
| 1414 | %tci%\end{document} |