New URL for NEMO forge!   http://forge.nemo-ocean.eu

Since March 2022 along with NEMO 4.2 release, the code development moved to a self-hosted GitLab.
This present forge is now archived and remained online for history.
Changeset 2282 for branches/nemo_v3_3_beta/DOC/TexFiles/Chapters/Chap_Model_Basics.tex – NEMO

Ignore:
Timestamp:
2010-10-15T16:42:00+02:00 (14 years ago)
Author:
gm
Message:

ticket:#658 merge DOC of all the branches that form the v3.3 beta

File:
1 edited

Legend:

Unmodified
Added
Removed
  • branches/nemo_v3_3_beta/DOC/TexFiles/Chapters/Chap_Model_Basics.tex

    r1224 r2282  
    77\minitoc 
    88 
     9 
     10\newpage 
     11$\ $\newline    % force a new ligne 
    912 
    1013% ================================================================ 
     
    164167 
    165168 
     169\newpage 
     170$\ $\newline    % force a new ligne 
     171 
    166172% ================================================================ 
    167173% The Horizontal Pressure Gradient 
     
    196202 the sea surface is the surface $z=0$. This well known approximation increases the surface  
    197203 wave speed to infinity and modifies certain other longwave dynamics ($e.g.$ barotropic  
    198  Rossby or planetary waves). In the present release of \NEMO, both strategies are still available.  
    199  They are further described in the next two sub-sections. 
     204 Rossby or planetary waves). The rigid-lid hypothesis is an obsolescent feature in modern  
     205 OGCMs. It has been available until the release 3.1 of  \NEMO, and it has been removed 
     206 in release 3.2 and followings. Only the free surface formulation is now described in the 
     207 this document (see the next sub-section). 
    200208 
    201209% ------------------------------------------------------------------------------------------------------------- 
     
    221229short with respect to the other processes described by the primitive equations. 
    222230 
    223 Three choices can be made regarding the implementation of the free surface in the model,  
     231Two choices can be made regarding the implementation of the free surface in the model,  
    224232depending on the physical processes of interest.  
    225233 
     
    238246of heat and salt contents is sufficient for the problem solved, then it is  
    239247sufficient to solve a linearized version of (\ref{Eq_PE_ssh}), which still allows  
    240 to take into account freshwater fluxes applied at the ocean surface \citep{Roullet2000}. 
    241  
    242 $\bullet$ For process studies not involving external waves nor surface freshwater  
    243 fluxes, it is possible to use the rigid lid approximation see (next  
    244 section). The ocean surface is then considered as a fixed surface, so that all  
    245 external waves are removed from the system.  
     248to take into account freshwater fluxes applied at the ocean surface \citep{Roullet_Madec_JGR00}. 
    246249 
    247250The filtering of EGWs in models with a free surface is usually a matter of discretisation  
    248 of the temporal derivatives, using the time splitting method \citep{Killworth1991, Zhang1992}  
     251of the temporal derivatives, using the time splitting method \citep{Killworth_al_JPO91, Zhang_Endoh_JGR92}  
    249252or the implicit scheme \citep{Dukowicz1994}. In \NEMO, we use a slightly different approach  
    250 developed by \citet{Roullet2000}: the damping of EGWs is ensured by introducing an  
     253developed by \citet{Roullet_Madec_JGR00}: the damping of EGWs is ensured by introducing an  
    251254additional force in the momentum equation. \eqref{Eq_PE_dyn} becomes:  
    252255\begin{equation} \label{Eq_PE_flt} 
     
    266269$i.e.$ they are stationary and damped. The diffusion regime applies to the modes shorter than  
    267270$T_c$. For longer ones, the diffusion term vanishes. Hence, the temporally unresolved EGWs  
    268 can be damped by choosing $T_c > \Delta t$. \citet{Roullet2000} demonstrate that  
     271can be damped by choosing $T_c > \rdt$. \citet{Roullet_Madec_JGR00} demonstrate that  
    269272(\ref{Eq_PE_flt}) can be integrated with a leap frog scheme except the additional term which  
    270273has to be computed implicitly. This is not surprising since the use of a large time step has a  
     
    272275Firstly, the damping of EGWs can be quantified through the magnitude of the additional term.  
    273276Secondly, the numerical scheme does not need any tuning. Numerical stability is ensured as  
    274 soon as $T_c > \Delta t$. 
     277soon as $T_c > \rdt$. 
    275278 
    276279When the variations of free surface elevation are small compared to the thickness of the first  
    277280model layer, the free surface equation (\ref{Eq_PE_ssh}) can be linearized. As emphasized  
    278 by \citet{Roullet2000} the linearization of (\ref{Eq_PE_ssh}) has consequences on the  
     281by \citet{Roullet_Madec_JGR00} the linearization of (\ref{Eq_PE_ssh}) has consequences on the  
    279282conservation of salt in the model. With the nonlinear free surface equation, the time evolution  
    280283of the total salt content is  
     
    287290integral over the free surface) vanishes when the nonlinear equation (\ref{Eq_PE_ssh})  
    288291is satisfied, so that the salt is perfectly conserved. When the free surface equation is  
    289 linearized, \citet{Roullet2000} show that the total salt content integrated in the fixed  
     292linearized, \citet{Roullet_Madec_JGR00} show that the total salt content integrated in the fixed  
    290293volume $D$ (bounded by the surface $z=0$) is no longer conserved: 
    291294\begin{equation} \label{Eq_PE_salt_content_linear} 
     
    295298 
    296299The right hand side of (\ref{Eq_PE_salt_content_linear}) is small in equilibrium solutions  
    297 \citep{Roullet2000}. It can be significant when the freshwater forcing is not balanced and  
     300\citep{Roullet_Madec_JGR00}. It can be significant when the freshwater forcing is not balanced and  
    298301the globally averaged free surface is drifting. An increase in sea surface height \textit{$\eta $}  
    299302results in a decrease of the salinity in the fixed volume $D$. Even in that case though,  
     
    309312its variations are driven by correlations of the time variation of surface salinity with the  
    310313sea surface height, which is a negligible term. This situation contrasts with the case of  
    311 the rigid lid approximation (following section) in which case freshwater forcing is  
    312 represented by a virtual salt flux, leading to a spurious source of salt at the ocean  
    313 surface \citep{Roullet2000}. 
    314  
    315 % ------------------------------------------------------------------------------------------------------------- 
    316 % Rigid-Lid Formulation 
    317 % ------------------------------------------------------------------------------------------------------------- 
    318 \subsection{Rigid-Lid formulation} 
    319 \label{PE_rigid_lid} 
    320  
    321 With the rigid lid approximation, we assume that the ocean surface ($z=0$) is a rigid lid  
    322 on which a pressure $p_s$ is exerted. This implies that the vertical velocity at the surface  
    323 is equal to zero. From the continuity equation \eqref{Eq_PE_continuity} and the kinematic  
    324 condition at the bottom \eqref{Eq_PE_w_bbc} (no flux across the bottom), it can be shown  
    325 that the vertically integrated flow $H{\rm {\bf \bar {U}}}_h$ is non-divergent (where the  
    326 overbar indicates a vertical average over the whole water column, i.e. from $z=-H$,  
    327 the ocean bottom, to $z=0$ , the rigid-lid). Thus, $\rm {\bf \bar {U}}_h$ can be derived  
    328 from a volume transport streamfunction $\psi$: 
    329 \begin{equation} \label{Eq_PE_u_psi} 
    330 \overline{\vect{U}}_h =\frac{1}{H}\left(   \vect{k} \times \nabla \psi   \right) 
    331 \end{equation} 
    332  
    333 As $p_s$ does not depend on depth, its horizontal gradient is obtained by forming the  
    334 vertical average of \eqref{Eq_PE_dyn} and using \eqref{Eq_PE_u_psi}: 
    335  
    336 \begin{equation} \label{Eq_PE_u_barotrope} 
    337 \frac{1}{\rho _o }\nabla _h p_s  
    338 =\overline{\vect{M}} -\frac{\partial \overline{\vect{U}} _h }{\partial t} 
    339 =\overline{\vect{M}}  
    340 -\frac{1}{H} \left[   \vect{k} \times \nabla \left( \frac{\partial \psi}{\partial t} \right)   \right] 
    341 \end{equation} 
    342  
    343 Here ${\rm {\bf M}} = \left( M_u,M_v \right)$ represents the collected contributions of the  
    344 Coriolis, hydrostatic pressure gradient, nonlinear and viscous terms in \eqref{Eq_PE_dyn}.  
    345 The time derivative of $\psi $ is the solution of an elliptic equation which is obtained from  
    346 the vertical component of the curl of (\ref{Eq_PE_u_barotrope}): 
    347 \begin{equation} \label{Eq_PE_psi} 
    348 \left[   {\nabla \times \left[ {\frac{1}{H} \vect{\bf k}  
    349   \times \nabla \left(   {\frac{\partial \psi }{\partial t}} \right)}   \right]} \; \right]_z  
    350 =\left[   {\nabla \times \overline{\vect{M}} }   \right]_z  
    351 \end{equation} 
    352  
    353 Using the proper boundary conditions, \eqref{Eq_PE_psi} can be solved to find $\partial_t \psi$  
    354 and thus using \eqref{Eq_PE_u_barotrope} the horizontal surface pressure gradient.  
    355 It should be noted that $p_s$ can be computed by taking the divergence of  
    356 \eqref{Eq_PE_u_barotrope} and solving the resulting elliptic equation. Thus the surface  
    357 pressure is a diagnostic quantity that can be recovered for analysis purposes. 
    358  
    359 A difficulty lies in the determination of the boundary condition on $\partial_t \psi$.  
    360 The boundary condition on velocity is that there is no flow normal to a solid wall,  
    361 $i.e.$ the coastlines are streamlines. Therefore \eqref{Eq_PE_psi} is solved with  
    362 the following Dirichlet boundary condition: $\partial_t \psi$ is constant along each  
    363 coastline of the same continent or of the same island. When all the coastlines are  
    364 connected (there are no islands), the constant value of $\partial_t \psi$ along the  
    365 coast can be arbitrarily chosen to be zero. When islands are present in the domain,  
    366 the value of the barotropic streamfunction will generally be different for each island  
    367 and for the continent, and will vary with respect to time. So the boundary condition is:  
    368 $\psi=0$ along the continent and $\psi=\mu_n$ along island $n$ ($1 \leq n \leq Q$),  
    369 where $Q$ is the number of islands present in the domain and $\mu_n$ is a time  
    370 dependent variable. A time evolution equation of the unknown $\mu_n$ can be found  
    371 by evaluating the circulation of the time derivative of the vertical average (barotropic)  
    372 velocity field along a closed contour around each island. Since the circulation of a  
    373 gradient field along a closed contour is zero, from \eqref{Eq_PE_u_barotrope} we have: 
    374 \begin{equation} \label{Eq_PE_isl_circulation} 
    375 \oint_n {\frac{1}{H}\left[ {{\rm {\bf k}}\times \nabla \left(  
    376 {\frac{\partial \psi }{\partial t}} \right)} \right] \cdot {\rm {\bf d}}\ell }  
    377 = \oint_n {\overline {\rm {\bf M}} \cdot {\rm {\bf d}}\ell }  
    378 \qquad  1 \leq n \leq Q 
    379 \end{equation} 
    380  
    381 Since (\ref{Eq_PE_psi}) is linear, its solution \textit{$\psi $} can be decomposed  
    382 as follows: 
    383 \begin{equation} \label{Eq_PE_psi_isl} 
    384 \psi =\psi _o +\sum\limits_{n=1}^{n=Q} {\mu _n \psi _n }  
    385 \end{equation} 
    386 where $\psi _o$ is the solution of \eqref{Eq_PE_psi} with $\psi _o=0$ long all  
    387 the coastlines, and where $\psi _n$ is the solution of \eqref{Eq_PE_psi} with  
    388 the right-hand side equal to $0$, and with $\psi _n =1$ long the island $n$,  
    389 $\psi _n =0$ along the other boundaries. The function $\psi _n$ is thus  
    390 independent of time. Introducing \eqref{Eq_PE_psi_isl} into  
    391 \eqref{Eq_PE_isl_circulation} yields: 
    392 \begin{multline} \label{Eq_PE_psi_isl_circulation} 
    393 \left[ {\oint_n {\frac{1}{H}  \left[ {{\rm {\bf k}}\times \nabla \psi _m } \right]\cdot {\rm {\bf d}}\ell } } \right]_{1\leq m\leqslant Q \atop 1\leq n\leqslant Q  } 
    394  \left( {\frac{\partial \mu _n }{\partial t}}  
    395 \right)_{1\leqslant n\leqslant Q}        \\ 
    396  =\left( {\oint_n {\left[ {\overline {\rm  
    397 {\bf M}} -\frac{1}{H}\left[ {{\rm {\bf k}}\times \nabla \left(  
    398 {\frac{\partial \psi _o }{\partial t}} \right)} \right]} \right]\cdot {\rm  
    399 {\bf d}}\ell } } \right)_{1\leqslant n\leqslant Q}  
    400 \end{multline} 
    401 which can be rewritten as: 
    402 \begin{equation} \label{Eq_PE_psi_isl_matrix} 
    403 {\rm {\bf A}}\;\left( {\frac{\partial \mu _n }{\partial t}}  
    404 \right)_{1\leqslant n\leqslant Q} ={\rm {\bf B}} 
    405 \end{equation} 
    406 where \textbf{A} is a $Q  \times Q$ matrix and \textbf{B} is a time dependent vector.  
    407 As \textbf{A} is independent of time, it can be calculated and inverted once. The time  
    408 derivative of the streamfunction when islands are present is thus given by: 
    409 \begin{equation} \label{Eq_PE_psi_isl_dt} 
    410 \frac{\partial \psi }{\partial t}=\frac{\partial \psi _o }{\partial  
    411 t}+\sum\limits_{n=1}^{n=Q} {{\rm {\bf A}}^{-1}{\rm {\bf B}}\;\psi _n }  
    412 \end{equation} 
    413  
    414  
     314the rigid lid approximation in which case freshwater forcing is represented by a virtual  
     315salt flux, leading to a spurious source of salt at the ocean surface  
     316\citep{Huang_JPO93, Roullet_Madec_JGR00}. 
     317 
     318\newpage 
     319$\ $\newline    % force a new ligne 
    415320 
    416321% ================================================================ 
     
    435340cannot be easily treated in a global model without filtering. A solution consists of introducing  
    436341an appropriate coordinate transformation that shifts the singular point onto land  
    437 \citep{MadecImb1996, Murray1996}. As a consequence, it is important to solve the primitive  
     342\citep{Madec_Imbard_CD96, Murray_JCP96}. As a consequence, it is important to solve the primitive  
    438343equations in various curvilinear coordinate systems. An efficient way of introducing an  
    439344appropriate coordinate transform can be found when using a tensorial formalism.  
     
    444349of the conservation laws of fluid dynamics. 
    445350 
    446 Let (\textbf{i},\textbf{j},\textbf{k}) be a set of orthogonal curvilinear coordinates on the sphere  
     351Let (\textit{i},\textit{j},\textit{k}) be a set of orthogonal curvilinear coordinates on the sphere  
    447352associated with the positively oriented orthogonal set of unit vectors (\textbf{i},\textbf{j},\textbf{k})  
    448353linked to the earth such that \textbf{k} is the local upward vector and (\textbf{i},\textbf{j}) are  
     
    682587term and can be viewed as a modification of the Coriolis parameter:  
    683588\begin{equation} \label{Eq_PE_cor+metric} 
    684 f \to f + \frac{1}{e_1 \; e_2}   \left(    v \frac{\partial e_2}{\partial i} 
    685                               -u \frac{\partial e_1}{\partial j}  \right) 
     589f \to f + \frac{1}{e_1\;e_2}  \left( v \frac{\partial e_2}{\partial i} 
     590                        -u \frac{\partial e_1}{\partial j}  \right) 
    686591\end{equation} 
    687592 
     
    690595the Coriolis parameter $f \to f+(u/a) \tan \varphi$. 
    691596 
    692 To sum up, the equations solved by the ocean model can be written in the following tensorial formalism: 
     597 
     598$\ $\newline    % force a new ligne 
     599 
     600To sum up, the curvilinear $z$-coordinate equations solved by the ocean model can be  
     601written in the following tensorial formalism: 
    693602 
    694603\vspace{+10pt} 
    695 $\bullet$ \textit{momentum equations} : 
    696  
    697 vector invariant form : 
     604$\bullet$ \textbf{Vector invariant form of the momentum equations} : 
     605 
    698606\begin{subequations} \label{Eq_PE_dyn_vect} 
    699 \begin{multline} \label{Eq_PE_dyn_vect_u} 
    700 \frac{\partial u}{\partial t}= 
    701    +   \left( {\zeta +f} \right)\,v                                     
    702    -   \frac{1}{2\,e_1} \frac{\partial}{\partial i} \left(  u^2+v^2   \right)  
    703    -   \frac{1}{e_3} w \frac{\partial u}{\partial k}       \\ 
    704    -   \frac{1}{e_1} \frac{\partial}{\partial i} \left( \frac{p_s+p_h }{\rho _o}    \right)     
    705 +   D_u^{\vect{U}}  +   F_u^{\vect{U}} 
    706 \end{multline} 
    707 \begin{multline} \label{Eq_PE_dyn_vect_v} 
    708 \frac{\partial v}{\partial t}= 
    709    -   \left( {\zeta +f} \right)\,u    
    710    -   \frac{1}{2\,e_2 }\frac{\partial }{\partial j}\left(  u^2+v^2  \right)        -   \frac{1}{e_3 }w\frac{\partial v}{\partial k}         \\ 
    711    -   \frac{1}{e_2 }\frac{\partial }{\partial j}\left( \frac{p_s+p_h }{\rho _o}  \right)     
    712 +  D_v^{\vect{U}}  +   F_v^{\vect{U}} 
    713 \end{multline} 
     607\begin{equation} \label{Eq_PE_dyn_vect_u} \begin{split} 
     608\frac{\partial u}{\partial t}  
     609= +   \left( {\zeta +f} \right)\,v                                     
     610   -   \frac{1}{2\,e_1}          \frac{\partial}{\partial i} \left(  u^2+v^2   \right)  
     611   -   \frac{1}{e_3    }  w     \frac{\partial u}{\partial k}      &      \\ 
     612   -   \frac{1}{e_1    }            \frac{\partial}{\partial i} \left( \frac{p_s+p_h }{\rho _o}    \right)     
     613   &+   D_u^{\vect{U}}  +   F_u^{\vect{U}}      \\ 
     614\\ 
     615\frac{\partial v}{\partial t} = 
     616       -   \left( {\zeta +f} \right)\,u    
     617       -   \frac{1}{2\,e_2 }        \frac{\partial }{\partial j}\left(  u^2+v^2  \right)    
     618       -   \frac{1}{e_3     }   w  \frac{\partial v}{\partial k}     &      \\ 
     619       -   \frac{1}{e_2     }        \frac{\partial }{\partial j}\left( \frac{p_s+p_h }{\rho _o}  \right)     
     620    &+  D_v^{\vect{U}}  +   F_v^{\vect{U}} 
     621\end{split} \end{equation} 
    714622\end{subequations} 
    715623 
    716 flux form: 
     624 
     625\vspace{+10pt} 
     626$\bullet$ \textbf{flux form of the momentum equations} : 
    717627\begin{subequations} \label{Eq_PE_dyn_flux} 
    718628\begin{multline} \label{Eq_PE_dyn_flux_u} 
     
    741651\end{multline} 
    742652\end{subequations} 
    743 where $\zeta$ is given by \eqref{Eq_PE_curl_Uh} and the surface pressure gradient formulation  
    744 depends on the one of the free surface: 
    745  
    746 $*$ free surface formulation 
    747 \begin{equation}\label{Eq_PE_dyn_sco} 
    748 \frac{1}{\rho _o }\nabla _h p_s =\left( {{\begin{array}{*{20}c} 
    749  {\frac{g}{\;e_1 }\frac{\partial \eta }{\partial i}} \hfill \\ 
    750  {\frac{g}{\;e_2 }\frac{\partial \eta }{\partial j}} \hfill \\ 
    751 \end{array} }} \right) 
    752 \qquad \text{where $\eta$ is solution of \eqref{Eq_PE_ssh} } 
    753 \end{equation} 
    754  
    755 $*$ rigid-lid approximation 
    756 \begin{equation}\label{Eq_PE_dyn_zco} 
    757 \frac{1}{\rho _o }\nabla _h p_s =\left( {{\begin{array}{*{20}c} 
    758  {\overline M _u +\frac{1}{H\;e_2 }\frac{\partial }{\partial j}\left(  
    759 {\frac{\partial \psi }{\partial t}} \right)}     \\ 
    760  {\overline M _v -\frac{1}{H\;e_1 }\frac{\partial }{\partial i}\left(  
    761 {\frac{\partial \psi }{\partial t}} \right)}        \\ 
    762 \end{array} }} \right) 
    763 \end{equation} 
    764 where ${\vect{M}}= \left( M_u,M_v \right)$ represents the collected contributions of nonlinear,  
    765 viscosity and hydrostatic pressure gradient terms in \eqref{Eq_PE_dyn_vect} and the overbar  
    766 indicates a vertical average over the whole water column ($i.e.$ from $z=-H$, the ocean bottom,  
    767 to $z=0$, the rigid-lid), and where the time derivative of $\psi$ is the solution of an elliptic equation: 
    768 \begin{multline} \label{Eq_psi_total} 
    769   \frac{\partial }{\partial i}\left[ {\frac{e_2 }{H\,e_1}\frac{\partial}{\partial i} 
    770                          \left( {\frac{\partial \psi }{\partial t}} \right)}   \right] 
    771 +\frac{\partial }{\partial j}\left[ {\frac{e_1 }{H\,e_2}\frac{\partial }{\partial j} 
    772                          \left( {\frac{\partial \psi }{\partial t}} \right)} \right] 
    773 = \\ 
    774   \frac{\partial }{\partial i}\left( {e_2 \overline M _v } \right) 
    775 - \frac{\partial }{\partial j}\left( {e_1 \overline M _u } \right) 
    776 \end{multline} 
     653where $\zeta$, the relative vorticity, is given by \eqref{Eq_PE_curl_Uh} and $p_s $,  
     654the surface pressure, is given by: 
     655\begin{equation} \label{Eq_PE_spg} 
     656p_s = \left\{ \begin{split}  
     657\rho \,g \,\eta &                                 \qquad  \qquad  \;   \qquad \text{ standard free surface} \\  
     658\rho \,g \,\eta &+ \rho_o \,\mu \,\frac{\partial \eta }{\partial t}      \qquad \text{ filtered     free surface}     
     659\end{split}  
     660\right. 
     661\end{equation} 
     662with $\eta$ is solution of \eqref{Eq_PE_ssh} 
    777663 
    778664The vertical velocity and the hydrostatic pressure are diagnosed from the following equations: 
     
    783669\frac{\partial p_h }{\partial k}=-\rho \;g\;e_3  
    784670\end{equation} 
    785  
    786671where the divergence of the horizontal velocity, $\chi$ is given by \eqref{Eq_PE_div_Uh}. 
    787672 
     
    809694in Chapter~\ref{SBC}. 
    810695 
     696 
    811697\newpage  
     698$\ $\newline    % force a new ligne 
    812699% ================================================================ 
    813 % Curvilinear z*-coordinate System 
     700% Curvilinear generalised vertical coordinate System 
    814701% ================================================================ 
    815 \section{Curvilinear \textit{z*}-coordinate System} 
    816  
    817 %>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    818 \begin{figure}[!b] \label{Fig_z_zstar}  \begin{center} 
    819 \includegraphics[width=1.0\textwidth]{./TexFiles/Figures/Fig_z_zstar.pdf} 
    820 \caption{(a) $z$-coordinate in linear free-surface case ; (b) $z-$coordinate in non-linear  
    821 free surface case (c) re-scaled height coordinate (become popular as the \textit{z*-}coordinate  
    822 \citep{Adcroft_Campin_OM04} ).} 
    823 \end{center}   \end{figure} 
    824 %>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
    825  
    826  
    827 In that case, the free surface equation is nonlinear, and the variations of volume are fully  
    828 taken into account. These coordinates systems is presented in a report \citep{Levier2007}  
    829 available on the \NEMO web site.  
    830  
    831 \gmcomment{ 
    832 The \textit{z*} coordinate approach is an unapproximated, non-linear free surface implementation  
    833 which allows one to deal with large amplitude free-surface 
    834 variations relative to the vertical resolution \citep{Adcroft_Campin_OM04}. In 
    835 the  \textit{z*} formulation, the variation of the column thickness due to sea-surface 
    836 undulations is not concentrated in the surface level, as in the $z$-coordinate formulation, 
    837 but is equally distributed over the full water column. Thus vertical 
    838 levels naturally follow sea-surface variations, with a linear attenuation with 
    839 depth, as illustrated by figure fig.1c . Note that with a flat bottom, such as in 
    840 fig.1c, the bottom-following  $z$ coordinate and  \textit{z*} are equivalent. 
    841 The definition and modified oceanic equations for the rescaled vertical coordinate 
    842  \textit{z*}, including the treatment of fresh-water flux at the surface, are 
    843 detailed in Adcroft and Campin (2004). The major points are summarized 
    844 here. The position ( \textit{z*}) and vertical discretization (\textit{z*}) are expressed as: 
    845  
    846 $H +  \textit{z*} = (H + z) / r$ and  $\delta \textit{z*} = \delta z / r$ with $r = \frac{H+\eta} {H}$ 
    847  
    848 Since the vertical displacement of the free surface is incorporated in the vertical 
    849 coordinate  \textit{z*}, the upper and lower boundaries are at fixed  \textit{z*} position,   
    850 $\textit{z*} = 0$ and  $\textit{z*} = ?H$ respectively. Also the divergence of the flow field  
    851 is no longer zero as shown by the continuity equation: 
    852  
    853 $\frac{\partial r}{\partial t} = \nabla_{\textit{z*}} \cdot \left( r \; \rm{\bf U}_h \right) 
    854       \left( r \; w\textit{*} \right) = 0 $ 
    855  
    856 } 
    857  
    858  
    859 \newpage  
    860 % ================================================================ 
    861 % Curvilinear s-coordinate System 
    862 % ================================================================ 
    863 \section{Curvilinear \textit{s}-coordinate System} 
    864 \label{PE_sco} 
    865  
    866 % ------------------------------------------------------------------------------------------------------------- 
    867 % Introduction 
    868 % ------------------------------------------------------------------------------------------------------------- 
    869 \subsection{Introduction} 
    870  
    871 Several important aspects of the ocean circulation are influenced by bottom topography.  
    872 Of course, the most important is that bottom topography determines deep ocean sub-basins,  
    873 barriers, sills and channels that strongly constrain the path of water masses, but more subtle  
    874 effects exist. For example, the topographic $\beta$-effect is usually larger than the planetary  
    875 one along continental slopes. Topographic Rossby waves can be excited and can interact  
    876 with the mean current. In the $z-$coordinate system presented in the previous section  
    877 (\S\ref{PE_zco}), $z-$surfaces are geopotential surfaces. The bottom topography is  
    878 discretised by steps. This often leads to a misrepresentation of a gradually sloping bottom  
    879 and to large localized depth gradients associated with large localized vertical velocities.  
    880 The response to such a velocity field often leads to numerical dispersion effects.  
    881 One solution to strongly reduce this error is to use a partial step representation of bottom  
    882 topography instead of a full step one \cite{Pacanowski_Gnanadesikan_MWR98}.  
    883 Another solution is to introduce a terrain-following coordinate system (hereafter $s-$coordinate)  
    884  
    885 The $s$-coordinate avoids the discretisation error in the depth field since the layers of  
    886 computation are gradually adjusted with depth to the ocean bottom. Relatively small  
    887 topographic features as well as  gentle, large-scale slopes of the sea floor in the deep  
    888 ocean, which would be ignored in typical $z$-model applications with the largest grid  
    889 spacing at greatest depths, can easily be represented (with relatively low vertical resolution).  
    890 A terrain-following model (hereafter $s-$model) also facilitates the modelling of the  
    891 boundary layer flows over a large depth range, which in the framework of the $z$-model  
    892 would require high vertical resolution over the whole depth range. Moreover, with a  
    893 $s$-coordinate it is possible, at least in principle, to have the bottom and the sea surface  
    894 as the only boundaries of the domain (nomore lateral boundary condition to specify).  
    895 Nevertheless, a $s$-coordinate also has its drawbacks. Perfectly adapted to a  
    896 homogeneous ocean, it has strong limitations as soon as stratification is introduced.  
    897 The main two problems come from the truncation error in the horizontal pressure  
    898 gradient and a possibly increased diapycnal diffusion. The horizontal pressure force  
    899 in $s$-coordinate consists of two terms (see Appendix~\ref{Apdx_A}), 
    900  
    901 \begin{equation} \label{Eq_PE_p_sco} 
    902 \left. {\nabla p} \right|_z =\left. {\nabla p} \right|_s -\frac{\partial  
    903 p}{\partial s}\left. {\nabla z} \right|_s  
    904 \end{equation} 
    905  
    906 The second term in \eqref{Eq_PE_p_sco} depends on the tilt of the coordinate surface  
    907 and introduces a truncation error that is not present in a $z$-model. In the special case  
    908 of a $\sigma-$coordinate (i.e. a depth-normalised coordinate system $\sigma = z/H$),  
    909 \citet{Haney1991} and \citet{Beckmann1993} have given estimates of the magnitude  
    910 of this truncation error. It depends on topographic slope, stratification, horizontal and  
    911 vertical resolution, the equation of state, and the finite difference scheme. This error  
    912 limits the possible topographic slopes that a model can handle at a given horizontal  
    913 and vertical resolution. This is a severe restriction for large-scale applications using  
    914 realistic bottom topography. The large-scale slopes require high horizontal resolution,  
    915 and the computational cost becomes prohibitive. This problem can be at least partially  
    916 overcome by mixing $s$-coordinate and step-like representation of bottom topography \citep{Gerdes1993a,Gerdes1993b,Madec1996}. However, the definition of the model  
    917 domain vertical coordinate becomes then a non-trivial thing for a realistic bottom  
    918 topography: a envelope topography is defined in $s$-coordinate on which a full or  
    919 partial step bottom topography is then applied in order to adjust the model depth to  
    920 the observed one (see \S\ref{DOM_zgr}. 
    921  
    922 For numerical reasons a minimum of diffusion is required along the coordinate surfaces  
    923 of any finite difference model. It causes spurious diapycnal mixing when coordinate  
    924 surfaces do not coincide with isoneutral surfaces. This is the case for a $z$-model as  
    925 well as for a $s$-model. However, density varies more strongly on $s-$surfaces than  
    926 on horizontal surfaces in regions of large topographic slopes, implying larger diapycnal  
    927 diffusion in a $s$-model than in a $z$-model. Whereas such a diapycnal diffusion in a  
    928 $z$-model tends to weaken horizontal density (pressure) gradients and thus the horizontal  
    929 circulation, it usually reinforces these gradients in a $s$-model, creating spurious circulation.  
    930 For example, imagine an isolated bump of topography in an ocean at rest with a horizontally  
    931 uniform stratification. Spurious diffusion along $s$-surfaces will induce a bump of isoneutral  
    932 surfaces over the topography, and thus will generate there a baroclinic eddy. In contrast,  
    933 the ocean will stay at rest in a $z$-model. As for the truncation error, the problem can be reduced by introducing the terrain-following coordinate below the strongly stratified portion of the water column  
    934 ($i.e.$ the main thermocline) \citep{Madec1996}. An alternate solution consists of rotating  
    935 the lateral diffusive tensor to geopotential or to isoneutral surfaces (see \S\ref{PE_ldf}.  
    936 Unfortunately, the slope of isoneutral surfaces relative to the $s$-surfaces can very large,  
    937 strongly exceeding the stability limit of such a operator when it is discretized (see Chapter~\ref{LDF}).  
    938  
    939 The $s-$coordinates introduced here \citep{Lott1990,Madec1996} differ mainly in two  
    940 aspects from similar models:  it allows  a representation of bottom topography with mixed  
    941 full or partial step-like/terrain following topography ; It also offers a completely general  
    942 transformation, $s=s(i,j,z)$ for the vertical coordinate. 
     702\section{Curvilinear generalised vertical coordinate System} 
     703\label{PE_gco} 
     704 
     705%\gmcomment{ 
     706The ocean domain presents a huge diversity of situation in the vertical. First the ocean surface is a time dependent surface (moving surface). Second the ocean floor depends on the geographical position, varying from more than 6,000 meters in abyssal trenches to zero at the coast. Last but not least, the ocean stratification exerts a strong barrier to vertical motions and mixing.  
     707Therefore, in order to represent the ocean with respect to the first point a space and time dependent vertical coordinate that follows the variation of the sea surface height $e.g.$ an $z$*-coordinate; for the second point, a space variation to fit the change of bottom topography $e.g.$ a terrain-following or $\sigma$-coordinate; and for the third point, one will be tempted to use a space and time dependent coordinate that follows the isopycnal surfaces, $e.g.$ an isopycnic coordinate. 
     708 
     709In order to satisfy two or more constrains one can even be tempted to mixed these coordinate systems, as in HYCOM (mixture of $z$-coordinate at the surface, isopycnic coordinate in the ocean interior and $\sigma$ at the ocean bottom) \citep{Chassignet_al_JPO03}  or OPA (mixture of $z$-coordinate in vicinity the surface and steep topography areas and $\sigma$-coordinate elsewhere) \citep{Madec_al_JPO96} among others. 
     710 
     711In fact one is totally free to choose any space and time vertical coordinate by introducing an arbitrary vertical coordinate : 
     712\begin{equation} \label{Eq_s} 
     713s=s(i,j,k,t) 
     714\end{equation} 
     715with the restriction that the above equation gives a single-valued monotonic relationship between $s$ and $k$, when $i$, $j$ and $t$ are held fixed. \eqref{Eq_s} is a transformation from the $(i,j,k,t)$ coordinate system with independent variables into the $(i,j,s,t)$ generalised coordinate system with $s$ depending on the other three variables through \eqref{Eq_s}. 
     716This so-called \textit{generalised vertical coordinate} \citep{Kasahara_MWR74} is in fact an Arbitrary Lagrangian--Eulerian (ALE) coordinate. Indeed, choosing an expression for $s$ is an arbitrary choice that determines which part of the vertical velocity (defined from a fixed referential) will cross the levels (Eulerian part) and which part will be used to move them (Lagrangian part). 
     717The coordinate is also sometime referenced as an adaptive coordinate \citep{Hofmeister_al_OM09}, since the coordinate system is adapted in the course of the simulation. Its most often used implementation is via an ALE algorithm, in which a pure lagrangian step is followed by regridding and remapping steps, the later step implicitly embedding the vertical advection \citep{Hirt_al_JCP74, Chassignet_al_JPO03, White_al_JCP09}. Here we follow the \citep{Kasahara_MWR74} strategy : a regridding step (an update of the vertical coordinate) followed by an eulerian step with an explicit computation of vertical advection relative to the moving s-surfaces. 
     718 
     719A key point here is that the $s$-coordinate depends on $(i,j)$ ==> horizontal pressure gradient... 
     720 
     721the generalized vertical coordinates used in ocean modelling are not orthogonal, which contrasts with many other applications in mathematical physics. Hence, it is useful to keep in mind the following properties that may seem odd on initial encounter. 
     722 
     723the horizontal velocity in ocean models measures motions in the horizontal plane, perpendicular to the local gravitational field. That is, horizontal velocity is mathematically the same regardless the vertical coordinate, be it geopotential, isopycnal, pressure, or terrain following. The key motivation for maintaining the same horizontal velocity component is that the hydrostatic and geostrophic balances are dominant in the large-scale ocean. Use of an alternative quasi-horizontal velocity, for example one oriented parallel to the generalized surface, would lead to unacceptable numerical errors. Correspondingly, the vertical direction is anti-parallel to the gravitational force in all of the coordinate systems. We do not choose the alternative of a quasi-vertical direction oriented normal to the surface of a constant generalized vertical coordinate.  
     724 
     725It is the method used to measure transport across the generalized vertical coordinate surfaces which differs between the vertical coordinate choices. That is, computation of the dia-surface velocity component represents the fundamental distinction between the various coordinates. In some models, such as geopotential, pressure,  
     726and terrain following, this transport is typically diagnosed from volume or mass conservation. In other models, such as isopycnal layered models, this transport is prescribed based on assumptions about the physical processes producing a flux across the layer interfaces.  
     727 
     728 
     729In this section we first establish the PE in the generalised vertical $s$-coordinate, then we discuss the particular cases available in \NEMO, namely $z$, $z$*, $s$, and $\tilde z$.   
     730%} 
    943731 
    944732% ------------------------------------------------------------------------------------------------------------- 
     
    1023811Add a few works on z and zps and s and underlies the differences between all of them 
    1024812\colorbox{yellow}{ $< = =$ end update}  } 
     813 
     814 
     815 
     816% ------------------------------------------------------------------------------------------------------------- 
     817% Curvilinear z*-coordinate System 
     818% ------------------------------------------------------------------------------------------------------------- 
     819\subsection{Curvilinear \textit{z*}--coordinate System} 
     820\label{PE_zco_star} 
     821 
     822%>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
     823\begin{figure}[!b] \label{Fig_z_zstar}  \begin{center} 
     824\includegraphics[width=1.0\textwidth]{./TexFiles/Figures/Fig_z_zstar.pdf} 
     825\caption{(a) $z$-coordinate in linear free-surface case ; (b) $z-$coordinate in non-linear  
     826free surface case (c) re-scaled height coordinate (become popular as the \textit{z*-}coordinate  
     827\citep{Adcroft_Campin_OM04} ).} 
     828\end{center}   \end{figure} 
     829%>>>>>>>>>>>>>>>>>>>>>>>>>>>> 
     830 
     831 
     832In that case, the free surface equation is nonlinear, and the variations of volume are fully  
     833taken into account. These coordinates systems is presented in a report \citep{Levier2007}  
     834available on the \NEMO web site.  
     835 
     836%\gmcomment{ 
     837The \textit{z*} coordinate approach is an unapproximated, non-linear free surface implementation  
     838which allows one to deal with large amplitude free-surface 
     839variations relative to the vertical resolution \citep{Adcroft_Campin_OM04}. In 
     840the  \textit{z*} formulation, the variation of the column thickness due to sea-surface 
     841undulations is not concentrated in the surface level, as in the $z$-coordinate formulation, 
     842but is equally distributed over the full water column. Thus vertical 
     843levels naturally follow sea-surface variations, with a linear attenuation with 
     844depth, as illustrated by figure fig.1c . Note that with a flat bottom, such as in 
     845fig.1c, the bottom-following  $z$ coordinate and  \textit{z*} are equivalent. 
     846The definition and modified oceanic equations for the rescaled vertical coordinate 
     847 \textit{z*}, including the treatment of fresh-water flux at the surface, are 
     848detailed in Adcroft and Campin (2004). The major points are summarized 
     849here. The position ( \textit{z*}) and vertical discretization (\textit{z*}) are expressed as: 
     850\begin{equation} \label{Eq_z-star} 
     851H +  \textit{z*} = (H + z) / r \quad \text{and} \ \delta \textit{z*} = \delta z / r \quad \text{with} \ r = \frac{H+\eta} {H} 
     852\end{equation}  
     853Since the vertical displacement of the free surface is incorporated in the vertical 
     854coordinate  \textit{z*}, the upper and lower boundaries are at fixed  \textit{z*} position,   
     855$\textit{z*} = 0$ and  $\textit{z*} = -H$ respectively. Also the divergence of the flow field  
     856is no longer zero as shown by the continuity equation: 
     857\begin{equation*}  
     858\frac{\partial r}{\partial t} = \nabla_{\textit{z*}} \cdot \left( r \; \rm{\bf U}_h \right) 
     859      \left( r \; w\textit{*} \right) = 0  
     860\end{equation*}  
     861%} 
     862 
     863 
     864% from MOM4p1 documentation 
     865 
     866To overcome problems with vanishing surface and/or bottom cells, we consider the  
     867zstar coordinate  
     868\begin{equation} \label{PE_} 
     869   z^\star = H \left( \frac{z-\eta}{H+\eta} \right) 
     870\end{equation} 
     871 
     872This coordinate is closely related to the "eta" coordinate used in many atmospheric  
     873models (see Black (1994) for a review of eta coordinate atmospheric models). It  
     874was originally used in ocean models by Stacey et al. (1995) for studies of tides  
     875next to shelves, and it has been recently promoted by Adcroft and Campin (2004)  
     876for global climate modelling. 
     877 
     878The surfaces of constant $z^\star$ are quasi-horizontal. Indeed, the $z^\star$ coordinate reduces to $z$ when $\eta$ is zero. In general, when noting the large differences between  
     879undulations of the bottom topography versus undulations in the surface height, it  
     880is clear that surfaces constant $z^\star$ are very similar to the depth surfaces. These properties greatly reduce difficulties of computing the horizontal pressure gradient relative to terrain following sigma models discussed in \S\ref{PE_sco}.  
     881Additionally, since $z^\star$ when $\eta = 0$, no flow is spontaneously generated in an  
     882unforced ocean starting from rest, regardless the bottom topography. This behaviour is in contrast to the case with "s"-models, where pressure gradient errors in  
     883the presence of nontrivial topographic variations can generate nontrivial spontaneous flow from a resting state, depending on the sophistication of the pressure  
     884gradient solver. The quasi-horizontal nature of the coordinate surfaces also facilitates the implementation of neutral physics parameterizations in $z^\star$ models using  
     885the same techniques as in $z$-models (see Chapters 13-16 of \cite{Griffies_Bk04}) for a  
     886discussion of neutral physics in $z$-models, as well as Section \S\ref{LDF_slp}  
     887in this document for treatment in \NEMO).  
     888 
     889The range over which $z^\star$ varies is time independent $-H \leq z^\star \leq 0$. Hence, all  
     890cells remain nonvanishing, so long as the surface height maintains $\eta > ?H$. This  
     891is a minor constraint relative to that encountered on the surface height when using  
     892$s = z$ or $s = z - \eta$.  
     893 
     894Because $z^\star$ has a time independent range, all grid cells have static increments  
     895ds, and the sum of the ver tical increments yields the time independent ocean  
     896depth %·k ds = H.  
     897The $z^\star$ coordinate is therefore invisible to undulations of the  
     898free surface, since it moves along with the free surface. This proper ty means that  
     899no spurious ver tical transpor t is induced across surfaces of constant $z^\star$ by the  
     900motion of external gravity waves. Such spurious transpor t can be a problem in  
     901z-models, especially those with tidal forcing. Quite generally, the time independent  
     902range for the $z^\star$ coordinate is a very convenient proper ty that allows for a nearly  
     903arbitrary ver tical resolution even in the presence of large amplitude fluctuations of  
     904the surface height, again so long as $\eta > -H$.  
     905 
     906%end MOM doc %%% 
     907 
     908 
     909 
     910\newpage  
     911% ------------------------------------------------------------------------------------------------------------- 
     912% Terrain following  coordinate System 
     913% ------------------------------------------------------------------------------------------------------------- 
     914\subsection{Curvilinear Terrain-following \textit{s}--coordinate} 
     915\label{PE_sco} 
     916 
     917% ------------------------------------------------------------------------------------------------------------- 
     918% Introduction 
     919% ------------------------------------------------------------------------------------------------------------- 
     920\subsubsection{Introduction} 
     921 
     922Several important aspects of the ocean circulation are influenced by bottom topography.  
     923Of course, the most important is that bottom topography determines deep ocean sub-basins,  
     924barriers, sills and channels that strongly constrain the path of water masses, but more subtle  
     925effects exist. For example, the topographic $\beta$-effect is usually larger than the planetary  
     926one along continental slopes. Topographic Rossby waves can be excited and can interact  
     927with the mean current. In the $z-$coordinate system presented in the previous section  
     928(\S\ref{PE_zco}), $z-$surfaces are geopotential surfaces. The bottom topography is  
     929discretised by steps. This often leads to a misrepresentation of a gradually sloping bottom  
     930and to large localized depth gradients associated with large localized vertical velocities.  
     931The response to such a velocity field often leads to numerical dispersion effects.  
     932One solution to strongly reduce this error is to use a partial step representation of bottom  
     933topography instead of a full step one \cite{Pacanowski_Gnanadesikan_MWR98}.  
     934Another solution is to introduce a terrain-following coordinate system (hereafter $s-$coordinate)  
     935 
     936The $s$-coordinate avoids the discretisation error in the depth field since the layers of  
     937computation are gradually adjusted with depth to the ocean bottom. Relatively small  
     938topographic features as well as  gentle, large-scale slopes of the sea floor in the deep  
     939ocean, which would be ignored in typical $z$-model applications with the largest grid  
     940spacing at greatest depths, can easily be represented (with relatively low vertical resolution).  
     941A terrain-following model (hereafter $s-$model) also facilitates the modelling of the  
     942boundary layer flows over a large depth range, which in the framework of the $z$-model  
     943would require high vertical resolution over the whole depth range. Moreover, with a  
     944$s$-coordinate it is possible, at least in principle, to have the bottom and the sea surface  
     945as the only boundaries of the domain (nomore lateral boundary condition to specify).  
     946Nevertheless, a $s$-coordinate also has its drawbacks. Perfectly adapted to a  
     947homogeneous ocean, it has strong limitations as soon as stratification is introduced.  
     948The main two problems come from the truncation error in the horizontal pressure  
     949gradient and a possibly increased diapycnal diffusion. The horizontal pressure force  
     950in $s$-coordinate consists of two terms (see Appendix~\ref{Apdx_A}), 
     951 
     952\begin{equation} \label{Eq_PE_p_sco} 
     953\left. {\nabla p} \right|_z =\left. {\nabla p} \right|_s -\frac{\partial  
     954p}{\partial s}\left. {\nabla z} \right|_s  
     955\end{equation} 
     956 
     957The second term in \eqref{Eq_PE_p_sco} depends on the tilt of the coordinate surface  
     958and introduces a truncation error that is not present in a $z$-model. In the special case  
     959of a $\sigma-$coordinate (i.e. a depth-normalised coordinate system $\sigma = z/H$),  
     960\citet{Haney1991} and \citet{Beckmann1993} have given estimates of the magnitude  
     961of this truncation error. It depends on topographic slope, stratification, horizontal and  
     962vertical resolution, the equation of state, and the finite difference scheme. This error  
     963limits the possible topographic slopes that a model can handle at a given horizontal  
     964and vertical resolution. This is a severe restriction for large-scale applications using  
     965realistic bottom topography. The large-scale slopes require high horizontal resolution,  
     966and the computational cost becomes prohibitive. This problem can be at least partially  
     967overcome by mixing $s$-coordinate and step-like representation of bottom topography \citep{Gerdes1993a,Gerdes1993b,Madec_al_JPO96}. However, the definition of the model  
     968domain vertical coordinate becomes then a non-trivial thing for a realistic bottom  
     969topography: a envelope topography is defined in $s$-coordinate on which a full or  
     970partial step bottom topography is then applied in order to adjust the model depth to  
     971the observed one (see \S\ref{DOM_zgr}. 
     972 
     973For numerical reasons a minimum of diffusion is required along the coordinate surfaces  
     974of any finite difference model. It causes spurious diapycnal mixing when coordinate  
     975surfaces do not coincide with isoneutral surfaces. This is the case for a $z$-model as  
     976well as for a $s$-model. However, density varies more strongly on $s-$surfaces than  
     977on horizontal surfaces in regions of large topographic slopes, implying larger diapycnal  
     978diffusion in a $s$-model than in a $z$-model. Whereas such a diapycnal diffusion in a  
     979$z$-model tends to weaken horizontal density (pressure) gradients and thus the horizontal  
     980circulation, it usually reinforces these gradients in a $s$-model, creating spurious circulation.  
     981For example, imagine an isolated bump of topography in an ocean at rest with a horizontally  
     982uniform stratification. Spurious diffusion along $s$-surfaces will induce a bump of isoneutral  
     983surfaces over the topography, and thus will generate there a baroclinic eddy. In contrast,  
     984the ocean will stay at rest in a $z$-model. As for the truncation error, the problem can be reduced by introducing the terrain-following coordinate below the strongly stratified portion of the water column  
     985($i.e.$ the main thermocline) \citep{Madec_al_JPO96}. An alternate solution consists of rotating  
     986the lateral diffusive tensor to geopotential or to isoneutral surfaces (see \S\ref{PE_ldf}.  
     987Unfortunately, the slope of isoneutral surfaces relative to the $s$-surfaces can very large,  
     988strongly exceeding the stability limit of such a operator when it is discretized (see Chapter~\ref{LDF}).  
     989 
     990The $s-$coordinates introduced here \citep{Lott_al_OM90,Madec_al_JPO96} differ mainly in two  
     991aspects from similar models:  it allows  a representation of bottom topography with mixed  
     992full or partial step-like/terrain following topography ; It also offers a completely general  
     993transformation, $s=s(i,j,z)$ for the vertical coordinate. 
     994 
     995 
     996\newpage  
     997% ------------------------------------------------------------------------------------------------------------- 
     998% Curvilinear z-tilde coordinate System 
     999% ------------------------------------------------------------------------------------------------------------- 
     1000\subsection{Curvilinear $\tilde{z}$--coordinate} 
     1001\label{PE_zco_tilde} 
     1002 
     1003 
    10251004 
    10261005\newpage  
Note: See TracChangeset for help on using the changeset viewer.