Subsections


Curvilinear generalised vertical coordinate System

The 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. Therefore, 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.

In 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) [Chassignet et al., 2003] or OPA (mixture of $ z$-coordinate in vicinity the surface and steep topography areas and $ \sigma$-coordinate elsewhere) [Madec et al., 1996] among others.

In fact one is totally free to choose any space and time vertical coordinate by introducing an arbitrary vertical coordinate :

$\displaystyle s=s(i,j,k,t)$ (2.20)

with the restriction that the above equation gives a single-valued monotonic relationship between $ s$ and $ k$, when $ i$, $ j$ and $ t$ are held fixed. (2.21) 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 (2.21). This so-called generalised vertical coordinate [Kasahara, 1974] 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). The coordinate is also sometime referenced as an adaptive coordinate [Hofmeister et al., 2009], 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 [Hirt et al., 1974, Chassignet et al., 2003, White et al., 2009]. Here we follow the [Kasahara, 1974] 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.

the 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.

The 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.

It 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, and 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.

In 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$.

The s-coordinate Formulation

Starting from the set of equations established in §2.3 for the special case $ k=z$ and thus $ e_3=1$, we introduce an arbitrary vertical coordinate $ s=s(i,j,k,t)$, which includes $ z$-, z*- and $ \sigma-$coordinates as special cases ($ s = z$, $ s=\textit{z*}$, and $ s=\sigma=z/H$ or $ =z/\left(H+\eta \right)$, resp.). A formal derivation of the transformed equations is given in Appendix A. Let us define the vertical scale factor by $ e_3=\partial_s z$ ($ e_3$ is now a function of $ (i,j,k,t)$ ), and the slopes in the (i,j) directions between $ s-$ and $ z-$surfaces by :

$\displaystyle \sigma _1 =\frac{1}{e_1 }\;\left. {\frac{\partial z}{\partial i}} \right\vert _s$   , and $\displaystyle \quad \sigma _2 =\frac{1}{e_2 }\;\left. {\frac{\partial z}{\partial j}} \right\vert _s$ (2.21)

We also introduce $ \omega$, a dia-surface velocity component, defined as the velocity relative to the moving $ s$-surfaces and normal to them:

$\displaystyle \omega = w - e_3   \frac{\partial s}{\partial t} - \sigma _1  u - \sigma _2  v  $ (2.22)

The equations solved by the ocean model (2.1) in $ s-$coordinate can be written as follows (see Appendix A.3):



$ \bullet$ Vector invariant form of the momentum equation :

\begin{multline}
\frac{\partial u }{\partial t}=
+ \left( {\zeta +f} \right) v...
...rho }{\rho _o}\sigma _1
+ D_u^{\vect{U}} + F_u^{\vect{U}} \quad
\end{multline}

\begin{multline}
\frac{\partial v }{\partial t}=
- \left( {\zeta +f} \right) u...
...ho }{\rho _o }\sigma _2
+ D_v^{\vect{U}} + F_v^{\vect{U}} \quad
\end{multline}



$ \bullet$ Vector invariant form of the momentum equation :

\begin{multline}
\frac{1}{e_3} \frac{\partial \left( e_3 u \right) }{\partial t...
...rho }{\rho _o}\sigma _1
+ D_u^{\vect{U}} + F_u^{\vect{U}} \quad
\end{multline}

\begin{multline}
\frac{1}{e_3} \frac{\partial \left( e_3 v \right) }{\partial t...
...ho }{\rho _o }\sigma _2
+ D_v^{\vect{U}} + F_v^{\vect{U}} \quad
\end{multline}

where the relative vorticity, $ \zeta$, the surface pressure gradient, and the hydrostatic pressure have the same expressions as in $ z$-coordinates although they do not represent exactly the same quantities. $ \omega$ is provided by the continuity equation (see Appendix A):

$\displaystyle \frac{\partial e_3}{\partial t} + e_3 \; \chi + \frac{\partial \omega }{\partial s} = 0$   with $\displaystyle \;\; \chi =\frac{1}{e_1 e_2 e_3 }\left[ {\frac{\partial \left( {e...
...}{\partial i}+\frac{\partial \left( {e_1 e_3  v} \right)}{\partial j}} \right]$ (2.23)



$ \bullet$ tracer equations:

\begin{multline}
\frac{1}{e_3} \frac{\partial \left( e_3 T \right) }{\partial t...
...rtial \left( {T \omega } \right)}{\partial k} + D^T + F^S \qquad
\end{multline}

\begin{multline}
\frac{1}{e_3} \frac{\partial \left( e_3 S \right) }{\partial t...
...rtial \left( {S \omega } \right)}{\partial k} + D^S + F^S \qquad
\end{multline}

The equation of state has the same expression as in $ z$-coordinate, and similar expressions are used for mixing and forcing terms.


Curvilinear z*-coordinate System

Figure 2.3: (a) $ z$-coordinate in linear free-surface case ; (b) $ z-$coordinate in non-linear free surface case ; (c) re-scaled height coordinate (become popular as the z*-coordinate [Adcroft and Campin, 2004] ).
\includegraphics[width=1.0\textwidth]{Fig_z_zstar}

In that case, the free surface equation is nonlinear, and the variations of volume are fully taken into account. These coordinates systems is presented in a report [Levier et al., 2007] available on the NEMO web site.

The z* coordinate approach is an unapproximated, non-linear free surface implementation which allows one to deal with large amplitude free-surface variations relative to the vertical resolution [Adcroft and Campin, 2004]. In the z* formulation, the variation of the column thickness due to sea-surface undulations is not concentrated in the surface level, as in the $ z$-coordinate formulation, but is equally distributed over the full water column. Thus vertical levels naturally follow sea-surface variations, with a linear attenuation with depth, as illustrated by figure fig.1c . Note that with a flat bottom, such as in fig.1c, the bottom-following $ z$ coordinate and z* are equivalent. The definition and modified oceanic equations for the rescaled vertical coordinate z*, including the treatment of fresh-water flux at the surface, are detailed in Adcroft and Campin (2004). The major points are summarized here. The position ( z*) and vertical discretization (z*) are expressed as:

$\displaystyle H + \textit{z*} = (H + z) / r$   and$\displaystyle  \delta \textit{z*} = \delta z / r$   with$\displaystyle  r = \frac{H+\eta} {H}$ (2.24)

Since the vertical displacement of the free surface is incorporated in the vertical coordinate z*, the upper and lower boundaries are at fixed z* position, $ \textit{z*} = 0$ and $ \textit{z*} = -H$ respectively. Also the divergence of the flow field is no longer zero as shown by the continuity equation:

$\displaystyle \frac{\partial r}{\partial t} = \nabla_{\textit{z*}} \cdot \left( r \; \rm {\bf U}_h \right) \left( r \; w\textit{*} \right) = 0$    

To overcome problems with vanishing surface and/or bottom cells, we consider the zstar coordinate

$\displaystyle z^\star = H \left( \frac{z-\eta}{H+\eta} \right)$ (2.25)

This coordinate is closely related to the "eta" coordinate used in many atmospheric models (see Black (1994) for a review of eta coordinate atmospheric models). It was originally used in ocean models by Stacey et al. (1995) for studies of tides next to shelves, and it has been recently promoted by Adcroft and Campin (2004) for global climate modelling.

The 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 undulations of the bottom topography versus undulations in the surface height, it is 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 §2.4.3. Additionally, since $ z^\star$ when $ \eta=0$, no flow is spontaneously generated in an unforced ocean starting from rest, regardless the bottom topography. This behaviour is in contrast to the case with "s"-models, where pressure gradient errors in the presence of nontrivial topographic variations can generate nontrivial spontaneous flow from a resting state, depending on the sophistication of the pressure gradient solver. The quasi-horizontal nature of the coordinate surfaces also facilitates the implementation of neutral physics parameterizations in $ z^\star$ models using the same techniques as in $ z$-models (see Chapters 13-16 of Griffies [2004]) for a discussion of neutral physics in $ z$-models, as well as Section §9.2 in this document for treatment in NEMO).

The range over which $ z^\star$ varies is time independent $ -H \leq z^\star \leq 0$. Hence, all cells remain nonvanishing, so long as the surface height maintains $ \eta > ?H$. This is a minor constraint relative to that encountered on the surface height when using $ s = z$ or $ s = z - \eta$.

Because $ z^\star$ has a time independent range, all grid cells have static increments ds, and the sum of the ver tical increments yields the time independent ocean depth The $ z^\star$ coordinate is therefore invisible to undulations of the free surface, since it moves along with the free surface. This proper ty means that no spurious ver tical transpor t is induced across surfaces of constant $ z^\star$ by the motion of external gravity waves. Such spurious transpor t can be a problem in z-models, especially those with tidal forcing. Quite generally, the time independent range for the $ z^\star$ coordinate is a very convenient proper ty that allows for a nearly arbitrary ver tical resolution even in the presence of large amplitude fluctuations of the surface height, again so long as $ \eta > -H$.


Curvilinear Terrain-following s-coordinate

Introduction

Several important aspects of the ocean circulation are influenced by bottom topography. Of course, the most important is that bottom topography determines deep ocean sub-basins, barriers, sills and channels that strongly constrain the path of water masses, but more subtle effects exist. For example, the topographic $ \beta$-effect is usually larger than the planetary one along continental slopes. Topographic Rossby waves can be excited and can interact with the mean current. In the $ z-$coordinate system presented in the previous section (§2.3), $ z-$surfaces are geopotential surfaces. The bottom topography is discretised by steps. This often leads to a misrepresentation of a gradually sloping bottom and to large localized depth gradients associated with large localized vertical velocities. The response to such a velocity field often leads to numerical dispersion effects. One solution to strongly reduce this error is to use a partial step representation of bottom topography instead of a full step one Pacanowski and Gnanadesikan [1998]. Another solution is to introduce a terrain-following coordinate system (hereafter $ s-$coordinate)

The $ s$-coordinate avoids the discretisation error in the depth field since the layers of computation are gradually adjusted with depth to the ocean bottom. Relatively small topographic features as well as gentle, large-scale slopes of the sea floor in the deep ocean, which would be ignored in typical $ z$-model applications with the largest grid spacing at greatest depths, can easily be represented (with relatively low vertical resolution). A terrain-following model (hereafter $ s-$model) also facilitates the modelling of the boundary layer flows over a large depth range, which in the framework of the $ z$-model would require high vertical resolution over the whole depth range. Moreover, with a $ s$-coordinate it is possible, at least in principle, to have the bottom and the sea surface as the only boundaries of the domain (nomore lateral boundary condition to specify). Nevertheless, a $ s$-coordinate also has its drawbacks. Perfectly adapted to a homogeneous ocean, it has strong limitations as soon as stratification is introduced. The main two problems come from the truncation error in the horizontal pressure gradient and a possibly increased diapycnal diffusion. The horizontal pressure force in $ s$-coordinate consists of two terms (see Appendix A),

$\displaystyle \left. {\nabla p} \right\vert _z =\left. {\nabla p} \right\vert _s -\frac{\partial p}{\partial s}\left. {\nabla z} \right\vert _s$ (2.26)

The second term in (2.33) depends on the tilt of the coordinate surface and introduces a truncation error that is not present in a $ z$-model. In the special case of a $ \sigma-$coordinate (i.e. a depth-normalised coordinate system $ \sigma = z/H$), Haney [1991] and Beckmann and Haidvogel [1993] have given estimates of the magnitude of this truncation error. It depends on topographic slope, stratification, horizontal and vertical resolution, the equation of state, and the finite difference scheme. This error limits the possible topographic slopes that a model can handle at a given horizontal and vertical resolution. This is a severe restriction for large-scale applications using realistic bottom topography. The large-scale slopes require high horizontal resolution, and the computational cost becomes prohibitive. This problem can be at least partially overcome by mixing $ s$-coordinate and step-like representation of bottom topography [Gerdes, 1993b, Gerdes, 1993a, Madec et al., 1996]. However, the definition of the model domain vertical coordinate becomes then a non-trivial thing for a realistic bottom topography: a envelope topography is defined in $ s$-coordinate on which a full or partial step bottom topography is then applied in order to adjust the model depth to the observed one (see §4.3.

For numerical reasons a minimum of diffusion is required along the coordinate surfaces of any finite difference model. It causes spurious diapycnal mixing when coordinate surfaces do not coincide with isoneutral surfaces. This is the case for a $ z$-model as well as for a $ s$-model. However, density varies more strongly on $ s-$surfaces than on horizontal surfaces in regions of large topographic slopes, implying larger diapycnal diffusion in a $ s$-model than in a $ z$-model. Whereas such a diapycnal diffusion in a $ z$-model tends to weaken horizontal density (pressure) gradients and thus the horizontal circulation, it usually reinforces these gradients in a $ s$-model, creating spurious circulation. For example, imagine an isolated bump of topography in an ocean at rest with a horizontally uniform stratification. Spurious diffusion along $ s$-surfaces will induce a bump of isoneutral surfaces over the topography, and thus will generate there a baroclinic eddy. In contrast, the ocean will stay at rest in a $ z$-model. As for the truncation error, the problem can be reduced by introducing the terrain-following coordinate below the strongly stratified portion of the water column ($ i.e.$ the main thermocline) [Madec et al., 1996]. An alternate solution consists of rotating the lateral diffusive tensor to geopotential or to isoneutral surfaces (see §2.5.2. Unfortunately, the slope of isoneutral surfaces relative to the $ s$-surfaces can very large, strongly exceeding the stability limit of such a operator when it is discretized (see Chapter 9).

The $ s-$coordinates introduced here [Madec et al., 1996, Lott et al., 1990] differ mainly in two aspects from similar models: it allows a representation of bottom topography with mixed full or partial step-like/terrain following topography ; It also offers a completely general transformation, $ s=s(i,j,z)$ for the vertical coordinate.


Curvilinear $ \tilde{z}$-coordinate

The $ \tilde{z}$-coordinate has been developed by Leclair and Madec [2011]. It is available in NEMO since the version 3.4. Nevertheless, it is currently not robust enough to be used in all possible configurations. Its use is therefore not recommended.

Gurvan Madec and the NEMO Team
NEMO European Consortium2017-02-17