MOM4 Technical Guide SEO
MOM4 Technical Guide SEO
It should be referenced as
A T ECHNICAL G UIDE TO MOM4
GFDL O CEAN G ROUP T ECHNICAL R EPORT N O . 5
S.M. Griffies, M.J. Harrison, R.C. Pacanowski, and A. Rosati
NOAA/Geophysical Fluid Dynamics Laboratory
Version prepared on September 18, 2003
Available online at www.gfdl.noaa.gov
Information about how to download and run MOM4 can be found at the GFDL
Flexible Modeling System (FMS) web site accessible from www.gfdl.noaa.gov.
This document was prepared using LATEX as described by Lamport (1994) and
Goosens et al. (1994).
CONTENTS
I Basics of MOM4 9
1 An introduction to MOM 13
1.1 What is MOM? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.2 Release of MOM4.0 October 2003 . . . . . . . . . . . . . . . . . . . . . 15
1.3 MOM4 documentation . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.4 Modeling frameworks . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.5 Some characteristics of MOM4 . . . . . . . . . . . . . . . . . . . . . . 17
1.6 Reproducing older results using MOM4.0 . . . . . . . . . . . . . . . . 22
1.7 Planned ocean model development . . . . . . . . . . . . . . . . . . . . 23
II Fundamentals of MOM4 37
16 Streamfunctions 185
16.1 Meridional-overturning streamfunction . . . . . . . . . . . . . . . . . 185
16.2 Vertically integrated transport . . . . . . . . . . . . . . . . . . . . . . . 190
21.1 The geometry of western (left) and eastern (right) open boundaries . 244
21.2 The geometry of southern (lower) and northern (upper) open bound-
aries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
Part I
Basics of MOM4
11
B ASICS OF MOM4
The purpose of this part of the MOM4 Guide is to familiarize the reader with
the basics of MOM4. There are two chapters, with the first providing an overview
of MOM4 and its relation to other versions of MOM. The second chapter provides
details of the computational aspects of MOM4. This part of the book should satisfy
those readers most interested in a quick overview and summary of MOM.
12
CHAPTER
ONE
An introduction to MOM
Contents
climate modelers have increased their collaboration with each other and with soft-
ware engineers and computational scientists. The aim of such collaborations is an
improved software infrastructure thus reducing the burden on any particular re-
searcher or group.
• coupler and data override: used to couple MOM4 to other component models
and/or datasets.
• initial and boundary data: regrids spherical fields to the generally non-spherical
ocean model grid
• grid and topography specification: sets model grid spacing and interpolates
spherical topography to the model grid
• diagnostic manager: to register and send fields to be written to a file for later
analysis
• field manager: for organizing multiple tracers for use especially in biogeo-
chemistry studies.
Being part of FMS greatly frees up those interested in developing physical and
numerical algorithms to focus on just that, instead of also needing to become ex-
perts in computational platforms and various software engineering issues. It also
allows for efficient input from computational scientists and engineers since they
can more readily focus on computational issues. Finally, it allows us in the ocean
modeling community to play a role in establishing common software infrastruc-
tures and coding standards, such as the ESMF efforts mentioned previously.
The FMS infrastructure was first released to the public earl 2002, with further
releases following on a regular basis. Notably, MOM4 represents the first major
model code to be released within FMS. Full documentation of FMS can be found at
www.gfdl.noaa.gov.
It is here that the researcher will find information about how to download and run
MOM4.
et al. (1991) and Griffies et al. (2001). After investigations leading to the Griffies
et al. (2001) paper, it was concluded that the explicit free surface method is most
suitable since it allows for efficient use of parallel computers while rendering the
model’s algorithms physically and numerically sound and simple. Hence, there is
only one external mode solver in MOM4, and it is a slight variant of the Griffies et
al. method detailed in Griffies (2004).
Other examples abound where numerous options were made available in MOM3
for physical parameterization schemes. Quite simply, MOM3 represented the end
of some ten years of research and experience with various approaches used in
MOM. Building up to that point required testing of numerous options prior to de-
ciding which ones to jettison. The advantage of this approach is that it allows for
ready examination of the many permutations and combinations leading to a well
tuned model. The disadvantage is that is leaves the inexperienced modeler with
little guidance since there are so many options.
In the development of MOM4, we attempted a balance between including mul-
tiple options and hard-line decisions about what would be supported and not sup-
ported. We hope that our choices will be suitable for a broad class of ocean climate
researchers.
/mom4/ocean_param/mixing/vert/kpp/ocean_vert_mix_coeff.F90
1.5. SOME CHARACTERISTICS OF MOM4 19
should be compiled. If instead, one wishes the constant vertical diffusivity ap-
proach, then
/mom4/ocean_param/mixing/vert/const/ocean_vert_mix_coeff.F90
should be compiled. This approach effectively replaces the selection of major ifdef
options with a pointer within a shell script to a desired module. The downside to
this approach is that some code in the different modules is similar, and so there is
a modest increase in code maintenance. The upside is that it cleans up the logic in
the separate modules.
• Multiple tracers are managed using the FMS field manager that organizes tracer
names, fluxes, sources, initializations, restarts, advection schemes, etc. This
manager was written, in particular, to serve the needs of ocean biogeochem-
istry research, as well as for use by atmospheric chemists.
• The FMS diagnostic manager is used to register and send fields to output for
analysis. The diagnostic manager and asociated diagnostic table allows for
the trivial addition of a new field to be added to the suite of model diagnostics
available for an experiment. This manager has been found to be extremely
useful and powerful.
• I/O is generally written in NetCDF. This capability includes files for restarts,
boundary forcing, initialization, topography, grids, sponges, etc. Some native
format capabilities remain, but with less support by GFDL than NetCDF.
20 CHAPTER 1. AN INTRODUCTION TO MOM
• The external mode solver is a variant of the Griffies et al. (2001) explicit free
surface. Top model grid cells have time dependent volume, thus allowing
for conservative fresh water input. Griffies (2004) presents full details and
rationale. There is no rigid-lid option in MOM4.
• Two equations of state are available. The most accurate is that described by
McDougall et al. (2003). In particular, the model’s density is a function of the
local potential temperature, salinity, and pressure. Pressure used for this cal-
culation is the time dependent hydrostatic pressure arising from fluid above
the point of interest, including the atmospheric pressure and pressure within
the ocean free surface (see Dewar et al. (1998)), and surface pressure is com-
puted using the local surface ocean density, not the constant Boussinesq den-
sity ρo . The second equation of state is a linearized equation for use in ideal-
ized Boussinesq models. Here, density is equated to potential density and is
a linear function of potential temperature. Nonlinear and pressure effects are
ignored in the linear equation of state. Chapter 20 details the MOM4 imple-
mentation.
available. The anisotropic scheme of Large et al. (2001) and Smith and McWilliams
(2002) has been implemented for both the Laplacian and biharmonic friction
operators, as has the associated Laplacian viscosity used in the Large et al.
(2001) paper. Griffies (2004) presents full details and rationale. Finally, the
methods of Holloway (1992) allow for horizontal friction to be computed as a
deviation from an approximate barotropic maximum entropy state (see Chap-
ter 23).
• Tracer advection is available using 2nd, 4th, 6th order centered schemes (Pacanowski
and Griffies (1999)), and the quicker scheme documented by Holland et al.
(1998) as well as Pacanowski and Griffies (1999). The algorithms for the 4th
and 6th order schemes assume constant grid spacing, thus simplifying their
code though compromising their accuracy on grids with large anisotropies.
MOM4 also provides for two multi-dimensional flux limited schemes ported
from the MIT GCM. These schemes are monotonic and quite efficient.
• The overflow scheme of Campin and Goosse (1999) has been implemented in
MOM4. This scheme has similar, yet complementary, characteristics relative
to the Beckmann and Döscher (1997) scheme. Chapter 15 describes the MOM4
implementation of Campin and Goosse (1999).
• Tidal forcing from eight lunar and solar constituents has been incorporated
into the free surface module of MOM4. Chapter 22 describes the implemen-
tation.
remain identical. Indeed, if such is not the case, then please carefully document the
differences and provide input to the developers.
• The leap-frog time stepping scheme has been the traditional approach for
time stepping the inviscid dynamics in MOM. Much of the code structure
assumes the leap-frog scheme, and such restricts ones ability to investigate
the utility of alternative approaches. The main problem with the leap-frog
scheme, at least for climate uses, is the inability to construct a discretely exact
conservation principle for tracers (e.g., Griffies et al. (2001) and Chapter 7).
This scheme can also exhibit substantial time-domain noise, even when using
the Robert-Asselin filter. Future development aims to remove the fundamen-
tal nature of the three-time level leap-frog scheme in MOM, so to allow testing
of alternative methods.
• MOM has always been discretized using a B-grid. With grid resolution in-
creasing, there are arguably some reasons for using the C-grid. In particular,
the C-grid provides far more flexibility in representing the complex topogra-
phy/geometry of an ocean basin since it allows for flow through single tracer
grid point channels. Work is planned to allow future GFDL models to use
either the B or C grids.
• MOM has always been a z-coordinate model. Recent advances based on the
isomorphism between depth and pressure suggest that minimal changes to
the dynamical kernal are needed to use sigma, eta, or pressure as the vertical
coordinate (e.g., Marshall et al. (2003) and Losch et al. (2003)). Such work is
planned for future development.
Contents
• grid type - grid locations, distances, metric terms and coriolis factors
• domain type - local and global indices, mpp domains flags and domain2d type.
26 CHAPTER 2. MOM4 QUICKSTART GUIDE
• time type - timestep counter, initial time, current time and time level indices.
• advective velocity type - horizontal and vertical advective velocities for tracers
and momentum.
• velocity type - horizonal velocity and density weighted velocity for non-boussinesq
calculations.
• tracer prog type - tracer concentration, surface and bottom fluxes and tracer
metadata including name and units. These are tracers that evolve via the
tracer equation.
• tracer diag type - tracer concentration and tracer metadata including name and
units. These are tracers that are diagnosed and so possess only a single time
level.
• external mode type - quantities related to the external mode, including surface
height and height tendency and freesurface forcing.
• ocean data type - Surface quantities for communication with fms coupler.
• ice ocean boundary type - surface flux quantities returned from the coupler.
Objects that are not contained in derived types include certain fields not associ-
ated with the ocean model as well as certain constants. Time independent objects
not contained in a derived type can be accessed by a module via the “use only”
statement. Time dependent objects are generally passed through subroutine inter-
faces. Maintaining this philosophy has reduced the head-aches associated with F90
predessor cycles (i.e., self-referential loops).
Coupler_main (driver)
Alternatively, the ocean model can be run in stand-alone mode using the stand-
alone driver (mom4/drivers/ocean solo.F90). Boundary fields are inserted into
the ice ocean boundary field using data override table entries in an identical fash-
ion to the coupled model. The stand-alone option does not require execution of
make xgrids and is appropriate for limited domain ocean configurations or more
idealized experiments.
src/preprocessing/generate_grids/ocean
src/preprocessing/generate_grids/ocean
An ASCII file containing index, depth pairs is prepared. edit grid makes the
necessary modifications and outputs to a separate grid specification file. Again,
refer to the script and code for this module to garner full details for the usage.
edit grid ASCII file format
src/preprocessing/mom4_prep/idealized_ic
src/preprocessing/regrid_3d
Notably, the regrid option can take a dataset that is originally gridded on a spherical
grid and regrid it to either a spherical ocean model grid, or to a tripolar ocean model
grid. Either form of the regridding uses a nearest neighbor approach.
As with initial conditions, it is necessary to setup boundary conditions on the
particular model grid (defined by the grid spec.nc file) used to run the experiment.
Boundary conditions can be either idealized, with code available in
2.6. STATIC MEMORY FOR OPTIMIZING ON SGI MACHINES 31
src/preprocessing/mom4_prep/idealized_ic
or taken from some dataset and regridded to the model grid, as done using files in
src/preprocessing/regrid_2d
The scripts and code are very similar to those used for the initial conditions, and
detailed documentation is provided in the code and scripts.
Note that when running a model using realistic boundary forcing, it is not nec-
essary to perform the regrid 2d step. An alternative is to leave the dataset on its
native grid and perform the regrid each time step of the model run. This is the
procedure for running a coupled model when the atmospheric grid differs from
the ocean. When the data has been regridded prior to running the ocean model, as
may be appropriate for an ocean-only run, then the “ongrid” flag can be set .true. in
the data override table, when running with the FMS coupler. Otherwise, the data
can reside on an arbitrary spherical grid which covers the ocean domain, in which
case “ongrid” is set false in the data override table.
Interior tracer restoring (or sponges) are generated exactly the same way as ini-
tial conditions, using regrid 3d.pl. An additional NetCDF file containing the damp-
ing timescale for the tracers should be generated as well. This can be accomplished
with an analysis package such as Ferret.
NOTE: missing values (such as land points) should be removed from the datasets before
using the regrid 2d and regrid 3d codes, or data override. Ferret provides tools for removing
missing values.
-Duse_libMPI
On the SGI Origin at GFDL, SHMEM is more efficient than MPI. When running
MOM4, the answers obtained with MPI agree at the bit-level to those obtained with
SHMEM.
may have been changed in the interim. All FMS modules write this information to
stdlog.
For the ocean model, namelist information is also written to stdout. In addition,
stdout contains various quality control statements highlighting what the model is
using, as well as warnings to indicate possible conflicts. MOM4.0 has a wide array
of options, and it is important that such verbosity be employed to communicate
to the user what the model is doing. Correspondingly, it is strongly encouraged
that the user frequently read printout files to ensure that the model is running in a
physically relevant manner. If users have suggestions for clearer or more complete
output statements, please provide such to the MOM4.0 developers.
• Those wishing to contribute new code to the main branch of mom4 should
document how the test cases are affected. If appropriate, a new test case
should be constructed exercising the new algorithm.
34 CHAPTER 2. MOM4 QUICKSTART GUIDE
There has been minimal tuning involved in the construction of test cases. Hence,
the researcher should not consider a test case as an “off the shelf model configura-
tion” appropriate for conducting relevant and publishable research. Indeed, there is
no guarantee that the test cases will run for an indefinite period of time. Providing
full support for “off the shelf” configurations is, unfortunately, beyond the abili-
ties of the MOM team. Given these caveats, the test cases nonetheless can provide
a useful starting point for the researcher who wishes to build models addressing
particular research questions.
2.9.2 Namelists
A particular experiment is defined by its grid, forcing, physics, and dynamics.
Much of the experimental details are specified by namelist settings. Namelists al-
low one to modify many of the experiment details while using the same executable,
and so not requiring a new compilation. This approach, introduced in MOM1,
greatly enhances the model flexibility and usability.
A complete listing of the namelist parameters available for a module can be
found via the associated html documentation provided with each module. Reading
the fortran source code is also, clearly, a way to understand what is available.
Fundamentals of MOM4
39
F UNDAMENTALS OF MOM4
The purpose of this part of the MOM4 Guide is to provide some grounding in
the fundamental physical, mathematical, and numerical aspects of MOM4. There
are many topics omitted here, with the book by Griffies (2004) suggested for those
wishing to understand fundamentals underlying ocean climate models. Nonethe-
less, there are many details outlined in this part of the MOM4 Guide that are useful
to those wishing to understand certain details of the model’s algorithms.
40
CHAPTER
THREE
Contents
x3 r
x2
x1
Figure 3.1: A schematic of the coordinates used for describing fluid dynamics on
a rotating sphere, where the rotation axis is aligned through the north pole of the
sphere. The coordinate 0 ≤ λ ≤ 2π is the longitude, with positive values measured
eastward from Greenwich, England. The coordinate φ is the latitude, with values
φ = 0 at the equator and φ = π /2(−π /2) at the north (south) poles. The radial dis-
tance r is measured here with respect to the center of the sphere. The explicit coor-
dinate transformations are x1 = r cos φ cos λ, x2 = r cos φ sin λ, and x3 = r sin φ.
Note that for many idealized geophysical fluid studies, cartesian coordinates refer
to those defined locally to a tangent plane at some point on the surface of the rotat-
ing sphere. Such β-plane or f -plane coordinates (e.g., Gill (1982); Pedlosky (1987))
are distinct from the cartesian coordinates defined here.
earth (see Figure 3.1). The use of orthogonal curvilinear coordinates (ξ 1 , ξ 2 ) allows
for the squared infinitesimal distance between two points in the ocean to be written
as
(ds)2 = (h1 dξ 1 )2 + (h2 dξ 2 )2 + dz2 , (3.1)
where the metric, or stretching functions, h1 and h2 are non-negative. In terms of
the dimensionful physical horizontal distances
dx = h1 dξ 1 (3.2)
2
dy = h2 dξ , (3.3)
the line element takes the compact form
(ds)2 = (dx)2 + (dy)2 + (dz)2 , (3.4)
the volume of an infinitesimal Eulerian region of the ocean is given by
dV = (h1 dξ 1 ) (h2 dξ 2 ) dz = dx dy dz, (3.5)
and the physical components of the horizontal partial derivatives are
∂ x = h−
1 ∂1
1
(3.6)
∂y = 2 ∂2 .
h− 1
(3.7)
Figure 3.2 illustrates these formulae. Notably, although the introduction of phys-
ical displacements brings the metric tensor into a form analogous to that for 3D-
Euclidean space, the non-Euclidean nature of the sphere manifests by the non-
vanishing commutator
∂ x , ∂ y = ∂ x ∂ y − ∂ y ∂ x = (∂ x ln dy) ∂ y − (∂ y ln dx) ∂ x ,
(3.8)
which vanishes only when the horizontal geometry is flat instead of curved. In
particular, the use of spherical coordinates leads to
tan φ
∂x , ∂ y = ∂x .
(3.9)
R
2
ξ
ξ
1
v = (u, w) (3.13)
M = v ∂ x ln dy − u ∂ y ln dx (3.14)
is the advection metric frequency arising from the non-Euclidean nature of the sphere.
The friction vector F(v) is associated with momentum transport due to molecular
3.2. OCEAN PRIMITIVE EQUATIONS 45
Coupled to the mass and momentum equations is the equation for conserva-
tion of tracers. There are three main ocean tracers: passive tracers, such as certain
biological constituents, the active tracers temperature and salinity, and dynamical
tracers such as potential vorticity. The terms passive and active refer to their influ-
ence on density, thus influencing dynamics through pressure. MOM4 time steps
the equations for passive and active tracers, whose form is generally given by
(ρ T ),t + ∇ · (ρ T v) = −∇ · (ρ F) + ρ S . (3.17)
In this equation, T is the tracer mass per mass of water (i.e., the tracer concentration),
The tracer flux F is interpreted as that arising from sub-grid-scale (SGS) molecular
processes, such as molecular diffusion. It will be reinterpreted in terms of ensemble
averages in Section 3.3. S is an interior tracer source term whose form depends on
details of the particular tracer.
∇·v = 0 (3.18)
(u)
u,t + ∇ · (v u) = −( f + M) ẑ ∧ u − ∇( p/ρo ) + F (3.19)
p,z = −ρ g (3.20)
T,t + ∇ · ( T v) = −∇ · F + S . (3.21)
The constraint ∇ · v = 0 means the Boussinesq fluid parcels conserve their volume
instead of their mass.
(ρ z D ),t = −ρo ∇ · U
e + ρw qw (3.23)
1 Ensemble averaging (Section 3.3) introduces more substantial SGS processes associated with
larger-scale turbulence, such as that occuring at the ocean’s microscale to mesoscale ranges.
46 CHAPTER 3. OCEAN PRIMITIVE EQUATIONS
There are two forms of fluid parcel kinematics of importance for ocean modeling:
parcels conserving their volume and parcels conserving their mass. In general,
parcel kinematic relations provide constraints on the fluid that are maintained re-
gardless the dynamics. Therefore, we believe it to be key to the integrity and usabil-
ity of the equations describing the ensemble averaged ocean, and consequently the
ocean model, that kinematics of the averaged fluid remains independent of dynam-
ical assumptions. In particular, we do not wish to require specification of unknown
closure terms, whose form depends on dynamical details, in order to determine
kinematic relations satisfied by the averaged ocean or the ocean model. Care in for-
mulating and interpreting the averaged equations is required to maintain this very
basic principle.
A key motivation for using density weighted averaging is that it assists in our
desire to keep the averaged parcel kinematics independent of dynamical closure as-
sumptions. Providing such a simple mapping between unaveraged and averaged
kinematics generally does not require much thought when averaging the Boussi-
nesq equations, since volume conservation ∇ · v = 0 is a linear constraint. Yet
for non-Boussinesq equations, mass conservation ρ,t + ∇ · (v ρ) = 0 is a nonlinear
constraint, thus requiring extra consideration.
Besides the mathematical utility of the density weighed approach for the non-
Boussinesq system, McDougall et al. (2002) argued for maintaining this average
even when considering the averaged Boussinesq equations. Their reasoning is
based on noting that the resulting Boussinesq system is far more accurate than the
mean-field equations resulting from non-density weighted averages. In this con-
text, accuracy is based on comparing with the small levels of diapycnal mixing in
the ocean interior.
48 CHAPTER 3. OCEAN PRIMITIVE EQUATIONS
ρ,t + ρo ∇ · v
e=0 (3.35)
e,t + ∇ · [(ρo /ρ) v
v eve] + (ρo /ρ) M
f ẑ ∧ v e − ∇( p/ρo ) + (ρ/ρo ) F(v)
e = −(ρ/ρo ) g ẑ − f ẑ ∧ v
(3.36)
(ρ T ),t + ρo ∇ · ( T v
e) = −∇ · (ρ F) + ρ S , (3.37)
hρi,t + ρo ∇ · hv ei = 0 (3.39)
z
D ∗ hρi e i + ρo q̃∗w .
= −ρo ∇ · hU (3.40)
,t
For the Boussinesq fluid, volume conservation for the parcel and column are given
by
∇ · hv
ei = 0 (3.41)
η∗,t = e i + q̃∗w .
−∇ · hU (3.42)
ρo q̃w = ρw qw . (3.44)
A starred quantity represents a modified mean field given by an infinite series of en-
semble mean correlations (Griffies (2004)). The precise relation between the starred
quantities and the correlations is not important, since the modified mean fields
provide the appropriate field to discretize in an ocean model, instead of the corre-
sponding ensemble mean field.
3.3. ENSEMBLE AVERAGED OCEAN PRIMITIVE EQUATIONS 49
The surface and bottom kinematic boundary conditions for the non-Boussinesq
fluid are
hρi η∗,t + ρo hu e i + ρo q̃∗w
e i · ∇η∗ = ρo hw at z = η∗ (3.45)
hue i · ∇ H + hw ei = 0 at z = − H (3.46)
and for the Boussinesq fluid we have
η∗,t + hu
e i · ∇η∗ = hwe i + q̃∗w at z = η∗ (3.47)
hu
e i · ∇ H + hw ei = 0 at z = − H. (3.48)
The tracer budget for the non-Boussinesq fluid is
(hρi h T iρ ),t + ρo ∇ · (hv
ei h T iρ ) = −ρo ∇ · hF̃sgs i + hρihSiρ (3.49)
and for the Boussinesq fluid
∂t h T iρ + ∇ · (hv
eih T iρ ) = −∇ · hF̃sgs i + hSiρ . (3.50)
Finally, the non-Boussinesq momentum budget is
ei,t + ∇ · (hviρ hv
hv f ẑ ∧ hviρ =
ei) + hMi
ei − ∇(h pi/ρo ) + hF̃vsgs i (3.51)
− (hρi/ρo ) g ẑ − f ẑ ∧ hv
whereas the Boussinesq budget is
hv
ei,t + ∇ · (hv
ei hv
ei) + hMi
f ẑ ∧ hv
ei =
ei − ∇(h pi/ρo ) + hF̃vsgs i. (3.52)
− (hρi/ρo ) g ẑ − f ẑ ∧ hv
ei = hviρ for a Boussinesq fluid, whereas ρo hv
Recall that hv ei = hρi hviρ for the
non-Boussinesq case. In the tracer equations, the SGS flux term is given by
ρ Fsgs = ρ Tρ0 vρ0 ≡ ρo F̃sgs , (3.53)
which dominates the contribution from molecular diffusion. The SGS friction vec-
tor F̃vsgs likewise incorporates SGS turbulence terms
(v) (v)
ρ Fsgs = ∇ · (ρ vρ0 vρ0 ) + ẑ ∧ ρ Mρ0 vρ0 ≡ ρo F̃sgs . (3.54)
The averaged equations have the same mathematical form as the unaveraged
equations given in Section 3.2. Precisely, the mapping between unaveraged and
averaged fields is given by
ρ → hρi (3.55)
p → h pi (3.56)
ρ
v → hvi (3.57)
e → hv
v ei (3.58)
ρ
T → hT i (3.59)
F̃ → hF̃sgs i (3.60)
ρ
S → hSi (3.61)
F̃v → hF̃vsgs i (3.62)
∗
η→η (3.63)
qw → q∗w . (3.64)
q̃w → q̃∗w . (3.65)
50 CHAPTER 3. OCEAN PRIMITIVE EQUATIONS
This mapping is very useful for purposes of analyzing properties of the two sys-
tems, such as their energetic balances discussed in Chapter 6. The fewer equations
we need to concern ourselves with, the better!
h pi → pmodel . (3.67)
hv
ei → vmodel . (3.68)
h T iρ → Tmodel . (3.69)
η∗ → ηmodel , (3.70)
ρ,t + ρo ∇ · v = 0 (3.72)
z
(h ρ ),t = −ρo ∇ · U + ρo qw (3.73)
v,t + ∇ · [(ρo /ρ) v v] + (ρo /ρ) M ẑ ∧ v =
− (ρ/ρo ) g ẑ − f ẑ ∧ v − ∇( p/ρo ) + (ρ/ρo ) F(v) (3.74)
(ρ T ),t + ρo ∇ · (v T ) = −ρo ∇ · F + ρ S , (3.75)
3.4. MAPPING TO OCEAN MODEL VARIABLES 51
These equations are combined with the surface and bottom kinematic boundary
conditions
ρ η,t + ρo u · ∇η = ρo w + ρo qw at z = η (3.79)
u · ∇H + w = 0 at z = − H. (3.80)
The mapping from unaveraged to averaged fields, and then from averaged to model
fields, is summarized in Table 3.1. This table is the key result from this chapter.
The continuous model equations presented above are identical in form to the
continuous unaveraged non-Boussinesq equations summarized in Section 3.2. Al-
though in the end somewhat trivial (i.e., what a round-about way to get back to the
same equations!), the intermediate steps reveal a nontrivial interpretation of the
fields discretized in the numerical model. It is hoped that such care in providing a
precise physical and mathematical interpretation for the variables adds rigor and
clarity to the foundations of MOM.
The Boussinesq model equations arise by setting ρmodel → ρo , except when
multiplying gravity. These equations are
∇·v = 0 (3.81)
η,t = −∇ · U + qw (3.82)
v,t + ∇ · (v v) + M ẑ ∧ v = −(ρ/ρo ) g ẑ − f ẑ ∧ v − ∇( p/ρo ) + F(v) (3.83)
T,t + ∇ · (v T ) = −∇ · F + S (3.84)
η,t + u · ∇η = w + qw at z = η (3.85)
u · ∇H + w = 0 at z = − H. (3.86)
As emphasized by McDougall et al. (2002) and Greatbatch et al. (2001), upon mak-
ing the hydrostatic approximation, these equations for the Boussinesq ocean model
are identical to those integrated by the Boussinesq version of MOM, with the ex-
ception of details that have been absorbed by the turbulence tracer and momentum
fluxes F and Fv . Additionally, McDougall et al. argue that the interpretation of
model fields as proposed here allows for the Boussinesq equations to be far more
accurate than the alternative interpretation. Hence, for this reason, and for reasons
of mathematical elegance, we prefer the interpretation summarized by Table 3.1 for
the variables carried by the Boussinesq and non-Boussinesq versions of MOM.
52 CHAPTER 3. OCEAN PRIMITIVE EQUATIONS
Contents
The purpose of this chapter is to detail the horizontal B-grid used in MOM4 as
well as the specification of field and grid values in halo regions.
B and C grids, as these are the two most commonly used grids in ocean modeling.
4.1. THE B-GRID USED IN MOM4 55
i U(i,j,k)
T(i,j,k)
Figure 4.1: Illustration of how fields are placed on the horizontal B-grid used in
MOM4. Velocity points Ui, j,k are placed to the northeast of tracer points Ti, j,k . Both
horizontal velocity components ui, j,k and vi, j,k are placed at the velocity point Ui, j,k .
and Ni, j . Figure 4.2 illustrates these points as oriented according to the tracer cell,
with Ti, j the usual tracer point. Ci, j lives at the northeast corner of the tracer cell,
and so represents B-grid velocity point Ui, j . Ni, j lives at the north face of the tracer
cell and so represents the C-grid meridional velocity point. Ei,k lives at the east
face of the tracer cell and so is the C-grid zonal velocity point. The geographical
coordinates of these four points is sufficient to place them on the discrete lattice.
N(i,.j)
C(i,j)
E(i,j)
T(i,j)
Figure 4.2: The four basic grid points for the B and C grids: Ti, j , Ci, j , Ei, j , and Ni, j .
Ti, j is the usual tracer point, with the corner Ci, j and side points Ei, j , Ni, j associated
with the tracer point.
56 CHAPTER 4. GRIDS AND HALOS
as well as the vertices of their corresponding grid cells is not sufficient. In addition,
we need information regarding the metric or stretching functions specific to the
coordinate system used to tile the sphere.
The traditional approach is to use spherical coordinates for tiling the sphere. In
this method, the distance between two points zonally displaced a finite distance
from one another is given by the analytic formula
Zλb
∆x[ a, b] = R cos φ dλ = ( R cos φ) (λb − λ a ), (4.1)
λa
and the distance between two points along a line of constant longitude is given by
Zφb
∆y[ a, b] = R dφ = R (φb − φ a ). (4.2)
φa
Writing this expression in a general manner leads to the generalized zonal and gen-
eralized meridional distance given by
( a)
ξ1
Z
∆x[ a, b] = h1 dξ1 (4.3)
(b)
ξ1
( a)
ξ2
Z
∆y[ a, b] = h2 dξ2 , (4.4)
(b)
ξ2
where (ξ1 , ξ2 ) represent generalized orthogonal coordinates, and (h1 , h2 ) are the
stretching functions specific to the coordinate system. They determine the distance
between two infinitesimally close points via the line element formula
With dx = h1 dξ1 and dy = h2 dξ2 , the line element formula takes the form of the
usual Cartesian expression
Given coordinates for the grid points and grid vertices, as well as the stretching
functions evaluated at these points, we can use the approximate expressions (4.8)
and (4.9) to compute distances between the T,U,N, and E points. Figure 4.3 shows
the notation for the grid distances that define four quarter-cells splitting up each
tracer and velocity cell. Shown is the notation used in the grid descriptor module
as well as that used in MOM4. The full dimensions of the tracer and velocity cells
are shown in Figure 4.4, where again the distances computed in the grid descriptor
module are translated into the grid distances used in MOM4. Finally, Figure 4.5
shows the distances specifying the separation between adjacent tracer and velocity
points.
• It maintains the usual spherical coordinate grid lines for latitudes southward
of the Arctic region, thus simplifying analysis.
• It is locally orthogonal, and so can be used with the MOM4 generalized hori-
zontal coordinates.
• Similar grids have been successfully run by the French OPA modeling group
and the Miami MICOM modeling group.
58 CHAPTER 4. GRIDS AND HALOS
ds_00_10_T ds_10_20_T
i due(i−1,j) duw(i,j)
U(i−1,j) U(i,j)
dtn(i,j) dus(i,j)
T(i,j)
dtw(i,j) dte(i,j)
dts(i,j) dun(i,j−1)
U(i,j−1)
Figure 4.3: Upper panel: Grid distances used to measure the distance between the
four fundamental grid points shown in Figure 4.2. These distances are computed
in the FMS grid descriptor module. The naming convention is based on a Cartesian
grid with the origin at the lower left corner of the tracer cell at (0, 0), the upper
right hand corner is (2, 2), the center at (1, 1), and all other points set accordingly.
The distances are then named as distances between these grid points. Note that
each tracer cell has a local Cartesian coordinate set as here, and so there is redu-
dancy in the various grid distances. Lower panel: When read into MOM4, the grid
distances set the distance between the tracer and velocity points used in the model
(Figure 4.1) and the sides of the corresponding grid cells. A translation of the upper
panel distance names to those used in MOM4 is made within MOM4’s ocean grid
module.
4.2. THE MURRAY (1996) TRIPOLAR GRID 59
ds_02_22_T
C(i,j)
(0,2) (1,2) (2,2)
ds_01_21_T T(i,j)
(0,1)
(1,1) (2,1)
ds_00_02_T ds_10_12_T ds_20_22_T
j
(0,0) (1,0) (2,0)
ds_00_20_T
i
dxtn(i,j) = dxue(i−1,j)
U(i−1,j) U(i,j)
T(i,j)
dxt(i,j) dyte(i,j)=dyun(i,j−1)
dyt(i,j)
U(i,j−1)
Figure 4.4: Grid cell distances used for computing the area of a grid cell. These
dimensions are related to the fundamental quarter-cell dimensions shown in Figure
4.3. Upper panel: distances computed in the FMS grid descriptor module. Lower
panel: names of the distances used in MOM4.
60 CHAPTER 4. GRIDS AND HALOS
T(i,j+1)
T(i+1,j+1)
ds_00_02_C C(i,j)
j
T(i,j) ds_00_20_C T(i+1,j)
T(i,j+1)
T(i+1,j+1)
dyu(i,j)
Figure 4.5: Distances between fundamental grid points (upper panel) as computed
by the grid descriptor module. These distances are taken into MOM4 and used to set
the distances between tracer and velocity points (lower panel).
4.2. THE MURRAY (1996) TRIPOLAR GRID 61
i
i=3ni/4
j=nj
i=0=ni i=ni/2
j
i=ni/4
i
Figure 4.6: Illustration of the grid lines forming the bipolar region in the Arctic.
This figure is taken after Figure 7 of Murray (1996). The thick outer boundary is a
line of constant latitude in the spherical coordinate grid. This latitude is typically at
the latitude nearest to 65◦ N. As in the spherical coordinate region, lines of constant
i move in a generalized eastward direction. They start from the bipolar south pole
at i = 0, which is identified with i = ni. The bipolar north pole is at i = ni /2. As
shown in Figure 4.7, the poles are centered at a velocity point. Lines of constant j
move in a generalized northward direction. The bipolar prime-meridion is situated
along the j-line with j = n j. This line defines the bipolar fold that bisects the tracer
grid. Its fold topology causes the velocity points centered along j = n j to have a
two-fold redundancy (see Figure 4.7 for more details).
62 CHAPTER 4. GRIDS AND HALOS
Figure 4.7: Schematic representation of the tracer and velocity cells on the bipolar
grid shown in Figure 4.6. The global computational domain consists of ni = 12
i-points for this example. The j = n j line bisects the tracer grid, which means there
are redundant velocity points along this line. Along an i −line of velocity points,
velocity cells with i = ni /2 live at the bipolar north pole, whereas velocity cells
with i = 0 = ni live at the bipolar south pole.
Figure 4.8: Schematic representation of fields living at the north and east faces of the
tracer cells as configured using the bipolar grid shown in Figure 4.6. Typical fields
of this sort are diffusive and advective tracer flux components, and so they are
components to a vector field, hence the vector notation. The global computational
domain consists of ni = 12 i-points for this example. The j = n j line bisects the
tracer grid, which means there are redundant velocity points along this line. Along
an i −line of velocity points, velocity cells with i = ni /2 live at the bipolar north
pole, whereas velocity cells with i = 0 = ni live at the bipolar south pole.
4.3. SPECIFYING FIELDS AND GRID DISTANCES WITHIN HALOS 63
PE(2)
halo for PE(0)
PE(1)
PE(0)
PE(3)
PE(5) PE(4)
Within the interior of the ocean model, away from global boundaries, the mapping
between domains is performed using an FMS utility that fills the halo points for
one local domain using information available to another local domain. Figure 4.9
illustrates this basic point. Shown is a central processor, arbitrarily labelled PE(0),
and a surrounding hatched region representing halo points. The width of the halo
is a function of the numerics used in the model. For second order numerics, a halo
width of a single point is sufficient. The values of fields and grid factors within the
halo are transmitted from the surrounding processors to PE(0) in order for PE(0) to
time step its portion of the ocean equations discretized on its local domain.
64 CHAPTER 4. GRIDS AND HALOS
For processors whose boundary touches the global model boundary, it is necessary
to specify whether the global boundary is a solid wall as in a sector model, peri-
odic as in a zonal channel, or folded as in the bipolar grid of Murray (1996). That
is, we must specify the model’s topology. Each of these three topologies requires
some special consideration, with the cases built into the MOM4 update boundary
condition module. Since these conditions are specific to the experimental design,
they are handled by a MOM4 module that sets the boundary conditions. We focus
here on the three common topologies supported by MOM4. A fourth case, open
boundary conditions, is discussed separately in chapter 21.
For a solid wall boundary condition, all fluxes passing across the walls are zeroed
out via masks, and fields within the solid wall are either trivial or masked. Hence,
no halo updates are necessary for fields and fluxes at solid walls. However, it is im-
portant to specify self-consistent grid distances separating points within the solid
wall from those within the model’s computational domain. The reason is that vari-
ous remapping operators require grid distances be well defined for all points within
the computational domain, including those distances reaching into the halo. See
Chapter 5 for details of remapping operators. For this reason, we extend the grid
into the solid wall halo so that resolution in this region is given by the resolution
between the two nearest interior points.
Zonally periodic channels (x-cyclic) are commonly run for idealized studies. Meri-
odionally periodic (y-cyclic) domains may also be of interest for simulations on an
f −plane or β-plane. For these reasons, we need to specify grid factors within the
halo assuming periodicity at the global domain boundary.
We focus here on the needs of the more common zonally periodic boundary
conditions, and refer to Figure 4.10. The same considerations hold for y-cyclic con-
ditions. For either case, we envision the grid wrapped onto itself in the appropriate
direction. With second order numerics, computation of the prognostic tracer in grid
cells Ti=1, j requires information regarding Ti=0, j . Likewise, Ti=ni, j requires informa-
tion about Ti=ni+1, j . Higher order numerics will need to reach out further.
First consider the eastern boundary of the domain where i = ni. For a single
grid halo, we need to specify values of fields living at the T, E, N, and C points at
i = ni + 1 (recall Figures 4.1 and 4.2 where the C point is equivalent to the B-grid U
point). Zonal periodicity renders the equalities
Figure 4.10: A zonally periodic array of tracer and velocity points with a single
halo point. In this example there are ni = 6 points in the global computational
domain, and halo = 1 point in the surrounding halo region. The cyclic mapping
leads us to specify halo points with values Ti=0, j = Ti=ni, j , Ti=ni+1, j = Ti=1, j , and
Ui=0, j = Ui=ni, j .
More generally, halo points with ni < i ≤ ni + halo acquire the x-cyclic mapping
At the western boundary, similar considerations lead to halo points 1 − halo ≤ i < 1
mapped to interior points according to
Likewise, scalars living at the northern face of a tracer cell contain a two-fold re-
dundancy of points along the j = n j line so that
For vector components living at U −points, such as the B-grid horizontal velocity
field, we associate transition across the j = n j meridion with a sign change
This sign change takes the right handed orientation into a right handed orientation
across the meridion. Likewise, for components of vector fluxes living at the north
face of a tracer cell, we have
Note that numerical roundoff may compromise these equalities in the model. Such
compromise will generally make the model energetics appear to be larger than
when running with the spherical grid, or with the tripolar grid with the fold closed
(debug tripolar = .true.).
Moving along a j-line, halo points for scalar fields with n j < j ≤ n j + halo are
evaluated according to the following rules
Ti, j = Tni−i+1,2n j− j+1
Ui, j = Uni−i,2n j− j
for n j < j ≤ n j + halo (4.27)
Ni, j = Nni−i+1,2n j− j
Ei, j = Eni−i,2n j− j+1
Vector components living at these points have the same index mapping along with
a sign flip for the field values.
Now consider the mappings needed to evaluate distances within halos. First
consider the distances associated with the tracer cells. By definition, dtei, j measures
4.3. SPECIFYING FIELDS AND GRID DISTANCES WITHIN HALOS 67
the distance between the tracer point Ti, j and its “eastern” neighbor Ei, j , and dtwi, j
is the distance between Ti, j with its “western” neighbor Ei−1, j , where “eastern” and
“western” are in a generalized sense. Mathematically, these distances are
∆x( Ti, j , Ei, j ) = dtei, j (4.32)
∆x( Ti, j , Ei−1, j ) = dtwi, j (4.33)
where ∆x( A, B) is the distance between points A and B computed according to
the generalized zonal distance in equation (4.8). The question is how to map these
distances across the bipolar fold. To do so, we note that if we are in a halo region
where n j < j ≤ n j + halo, then the scalar mappings given by equation (4.27) lead
to
∆x( Ti, j , Ei, j ) = ∆x( Tni−i+1,2n j− j+1 , Eni−i,2n j− j+1 ) (4.34)
∆x( Ti, j , Ei−1, j ) = ∆x( Tni−i+1,2n j− j+1 , Eni−i+1,2n j− j+1 ). (4.35)
Comparison of these equalities with the definitions of dte and dtw then leads to the
halo cell relations
dtei, j = dtwni−i+1,2n j− j+1
for n j < j ≤ n j + halo (4.36)
dtwi, j = dteni−i+1,2n j− j+1
Distances to the northern and southern faces of the tracer cell, dtn and dts, are
defined by
∆y( Ti, j , Ni, j ) = dtni, j (4.37)
∆y( Ti, j , Ni, j−1 ) = dtsi, j (4.38)
where ∆y is the generalized meridional distance given by equation (4.9). Equation
(4.27) indicate that within the halo region n j < j ≤ n j + halo,
∆y( Ti, j , Ni, j ) = ∆y( Tni−i+1,2n j− j+1 , Nni−i+1,2n j− j ) (4.39)
∆y( Ti, j , Ni, j−1 ) = ∆y( Tni−i+1,2n j− j+1 , Nni−i+1,2n j− j+1 ). (4.40)
Comparison of these equalities with the definitions of dtn and dts leads to the halo
cell relations
dtni, j = dtsni−i+1,2n j− j+1
for n j < j ≤ n j + halo (4.41)
dtsi, j = dtnni−i+1,2n j− j+1
Velocity cell distances are defined by
∆x(Ui, j , Ni+1, j ) = duei, j (4.42)
∆x(Ui, j , Ni, j ) = duwi, j (4.43)
∆y(Ui, j , Ei, j+1 ) = duni, j (4.44)
∆y(Ui, j , Ei, j ) = dusi, j (4.45)
Equation (4.27) indicate that within the halo region n j < j ≤ n j + halo,
∆x(Ui, j , Ni+1, j ) = ∆x(Uni−i,2n j− j , Nni−i,2n j− j ) (4.46)
∆x(Ui, j , Ni, j ) = ∆x(Uni−i,2n j− j , Nni−i+1,2n j− j ) (4.47)
∆y(Ui, j , Ei, j+1 ) = ∆x(Uni−i,2n j− j , Eni−i,2n j− j ) (4.48)
∆y(Ui, j , Ei, j ) = ∆x(Uni−i,2n j− j , Eni−i,2n j− j+1 ), (4.49)
68 CHAPTER 4. GRIDS AND HALOS
dus(ni/2+1,nj)
dtn(ni,nj) dun(1,nj) dtn(ni/2+1,nj) dus(ni/2,nj)
dun(ni/2,nj)
U(ni,nj) duw(ni,nj) U(ni/2+1,nj) due(ni/2,nj) U(ni/2,nj)
due(ni/2+1,nj) duw(ni/2+1,nj)
dtn(1,nj) dun(ni/2+1,nj)
dus(1,nj) dtn(ni/2,nj) dun(ni/2,nj) dus(ni/2,nj)
T(1,nj) T(ni/2,nj)
dtw(1,nj) dte(1,nj) dtw(ni/2,nj) dte(ni/2,nj)
Figure 4.11: Placement of quarter-cells distances at the bipolar fold. For this exam-
ple, there are ni = 4 points in the generalized zonal computational domain. Equiv-
alance of grid factors on the fold leads to the two-fold redundancy for velocity cell
distances duei,n j = duwni−i,n j and dusi,n j = dunni−i,n j .
Figures 4.12, 4.13, and 4.14 show these distances for regions surrounding the bipo-
lar fold. To generate the redudancy conditions and halo mappings, we again use
the scalar mappings given by equation (4.27). Using these relations we see that
redundancy is satisfied by the distances
Equation (4.27) indicates that within the halo region n j < j ≤ n j + halo,
dyt(ni,nj) dyt(ni/2+1,nj)
T(1,nj) T(ni/2,nj)
dxt(1,nj) dxt(ni/2,nj)
dyte(1,nj) dyte(ni/2,nj)
dyt(1,nj) dyt(ni/2,nj)
Figure 4.12: Placement of tracer cell dimensions at the bipolar fold. For this
example, there are ni = 4 points in the generalized zonal computational do-
main. Equivalance of grid factors on the fold leads to the two-fold redundancy
dxtni,n j = dxtnni−i+1,n j .
4.3. SPECIFYING FIELDS AND GRID DISTANCES WITHIN HALOS 71
T(ni,nj) T(ni/2+1,nj)
dyu(ni/2+1,nj)
dyu(1,nj)
T(1,nj) T(ni/2,nj)
dxte(ni/2+1,nj)
T(ni,nj) T(ni/2+1,nj)
T(1,nj) T(ni/2,nj)
dxte(1,nj)
Figure 4.14: Grid distances for tracer points at the bipolar fold.
72 CHAPTER 4. GRIDS AND HALOS
Table 4.1: Summary of the halo mappings and redundancies realized at the bipolar
fold. The symbol ε is 1 for scalar fields, and −1 for horizontal components of vector
fields.
CHAPTER
FIVE
Contents
and described in the Griffies et al. (2001) paper. This issue is relevant for the Boussi-
nesq and non-Boussinesq versions of MOM4. We visit this issue in Section 5.3.
where the distances dus and dun are lengths along sides of the four quarter-cells
comprising a single velocity cell (Figure 5.2). Likewise, the volume per unit length
per time passing across the meridional face through the velocity point Ui+1, j is
given by
Et(i + 1, j) dus(i + 1, j) + Et(i + 1, j + 1) dun(i + 1, j), (5.2)
and the volume per unit length per time passing across the eastern face of the ve-
locity cell Ui, j is given by
Eu(i, j) dytn(i + 1, j), (5.3)
where Eu is to be determined in terms of Et, and the grid distance dytn is the merid-
ional distance between tracer points, as defined in Figure 5.3.
76CHAPTER 5. ADVECTION VELOCITY COMPONENTS AND REMAPPING OPERATORS
U(i,j+1)
T(i+1,j+1)
T(i,j+1)
Et(i,j+1) Et(i+1,j+1)
U(i+1,j)
U(i,j) Eu(i,j)
T(i,j)
Et(i,j) T(i+1,j) Et(i+1,j)
the vertical flux of volume per unit length passing across the northern face of the
velocity cell Ui, j is given by
Bt(i, j + 1) dte(i, j + 1) + Bt(i + 1, j + 1) dtw(i + 1, j + 1), (5.9)
and the vertical flux of volume passing through the velocity cell is given by
Bu(i, j) dxu(i, j) dyu(i, j). (5.10)
Assuming that the total flux passing through the velocity cell is equivalent to that
passing across the northern plus southern parts of the cell leads to
dun(i,j)
duw(i,j) due(i,j)
U(i,j)
dtn(i,j)
dus(i,j)
T(i,j)
dtw(i,j) dte(i,j)
dts(i,j)
Figure 5.2: Time independent horizontal grid distances (meters) used for the tracer
cell Ti, j and velocity cell Ui, j in MOM4. These “quarter-cell” distances are refined
relative to those shown in Figures 5.4 and 5.5, and they are needed for the remap-
ping between T and U cells when computing advection velocities. All distances are
functions of both i and j due to the use of generalized orthogonal coordinates. Com-
paring with Figures 5.4 and 5.5 reveals the identities dtw(i, j) + dte(i, j) = dxt(i, j),
dts(i, j) + dtn(i, j) = dyt(i, j), duw(i, j) + due(i, j) = dxu(i, j), and dus(i, j) +
dun(i, j) = dyu(i, j).
5.3. REMAPPING OPERATOR FOR VERTICAL FLUXES 79
U(i,j+1)
dyun(i,j) T(i+1,j+1)
T(i,j+1)
dytn(i,j) dxue(i,j)
U(i+1,j)
U(i,j)
T(i,j)
dxte(i,j) T(i+1,j)
Figure 5.3: Time independent horizontal grid distances (meters) setting the spac-
ing between tracer and velocity points in MOM4. All distances are functions of
both i and j due to the use of generalized orthogonal coordinates. When these
distances are combined with those in Figures 5.4 and 5.5, and the quarter-cell dis-
tances given in Figure 5.2, we then have full information about the discrete hori-
zontal T and U cells on the model grid. Note there is some redundancy with the
distances defined in Figures 5.4 and 5.5, where we have dytn(i, j) = dyue(i − 1, j),
dxte(i, j) = dxun(i, j − 1), dxue(i, j) = dxtn(i + 1, j), and dyun(i, j) = dyte(i, j + 1).
Additionally, comparision with Figure 5.2 leads to the identities dyun(i, j) =
dun(i, j) + dus(i, j + 1), dxue(i, j) = due(i, j) + duw(i + 1, j), dytn(i, j) = dtn(i, j) +
dts(i, j + 1), and dxte(i, j) = dte(i, j) + dtw(i + 1, j).
80CHAPTER 5. ADVECTION VELOCITY COMPONENTS AND REMAPPING OPERATORS
dxtn(i,j)
dyt(i,j)
Figure 5.4: Time independent horizontal grid distances (meters) used for the tracer
cell Ti, j in MOM4. dxti, j and dyti, j are the grid distances of the tracer cell in the
generalized zonal and meridional directions, and dati, j = dxti, j dyti, j is the area of
the cell. The grid distance dxtni, j is the zonal width of the north face of a tracer cell,
and dytei, j is the meridional width of the east face. Note that the tracer point Ti, j is
not generally at the center of the tracer cell. Distances are functions of both i and j
due to the use of generalized orthogonal coordinates.
dxun(i,j)
dyu(i,j)
U(i,j)
dyue(i,j)
dxu(i,j)
Figure 5.5: Time independent horizontal grid distances (meters) used for the ve-
locity cell Ui, j in MOM4. dxui, j and dyui, j are the grid distances of the velocity cell
in the generalized zonal and meridional directions, and daui, j = dxui, j dyui, j is the
area of the cell. The grid distance dxuni, j is the zonal width of the north face of
a velocity cell, and dyuei, j is the meridional width of the east face. Note that the
velocity point Ui, j is not generally at the center of the velocity cell. Distances are
functions of both i and j due to the use of generalized orthogonal coordinates.
5.4. REMAPPING ERROR 81
MOM4 computes a diagnostic that examines the differences between the two ap-
proaches for computing vertical advective velocities. It reports the difference as
a “remapping error.” If the numerical discretization is self-consistent, then the
remapping error for a spherical grid will be roundoff, with values on the order
of 10−20 m s−1 common. Therefore, with a spherical grid, the remapping error pro-
vides a check on the self-consistency of the grid distances and the remapping op-
erators. Effectively, what is done is to check that volume is conserved with the
remapping operators. Even when running a non-Boussinesq model, the remapping
operators are constructed to respect volume conservation.
For a grid defined via a nonlinear transformation of the spherical grid, such as the
bipolar region of the tripolar grid, the grid no longer maintains a linear relation
between tracer and velocity cell distances. The result is a nontrivial remapping
error. This error can be reduced by defining new remapping operators that account
for a generally nonlinear relation between tracer and velocity grid distances. Such
remains to be done for MOM4.
One consequence of the nonzero remapping error is that for a flat bottom model
in regions where the grid distances are nonlinearly related, w bui, j,k= Nk does not
vanish, even though continuity is maintained for all the grid cells. The problem is
that w bui, j,k=0 is defined by
Contents
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
6.1.1 The utility of discrete energy conversions . . . . . . . . . . 84
6.1.2 Continuous equations for the ocean model . . . . . . . . . 85
6.1.3 Kinetic energy budget for a continuum ocean model parcel 85
6.1.4 Semi-discrete momentum budget . . . . . . . . . . . . . . 86
6.1.5 Vertical advective velocities . . . . . . . . . . . . . . . . . . 86
6.2 Pressure work conversions . . . . . . . . . . . . . . . . . . . . . . 87
6.2.1 Continuum results: Part I . . . . . . . . . . . . . . . . . . . 87
6.2.2 Continuum results: Part II . . . . . . . . . . . . . . . . . . . 88
6.2.3 B-grid results . . . . . . . . . . . . . . . . . . . . . . . . . . 89
6.2.3.1 Defining the horizontal advection velocities . . . 90
6.2.3.2 Completing the manipulations for P1 . . . . . . . 91
6.2.3.3 Correspondence to the continuum results . . . . 92
6.2.3.4 The sigma-correction term P2 . . . . . . . . . . . 93
6.2.3.5 Summary . . . . . . . . . . . . . . . . . . . . . . . 94
6.3 Kinetic energy advection . . . . . . . . . . . . . . . . . . . . . . . 95
6.3.1 Continuum results . . . . . . . . . . . . . . . . . . . . . . . 95
6.3.2 B-grid results . . . . . . . . . . . . . . . . . . . . . . . . . . 96
6.4 Kinetic energy in the external and internal modes . . . . . . . . 99
6.5 A caveat regarding the tripolar grid . . . . . . . . . . . . . . . . . 100
6.1 Introduction
This chapter is concerned with details of how the discrete model respects the con-
version between various forms of energy. For example, how does work done by
currents against the horizontal pressure gradient get converted into work against
gravity and/or compression? How does globally integrated discrete advection of
momentum get converted to boundary contributions? How do conversions differ
for Boussinesq and non-Boussinesq fluids?
As discussed in Griffies (2004), energy conversions in the continuum largely
follow from manipulations using the kinematic and dynamic balances. Analogous
manipulations occur on the lattice, yet with more care given to how various terms
are discretized. Maintaining exact discrete energic conversions is neither neces-
sary nor sufficient for ensuring a physically realistic solution. However, without
analytical solutions to compare with, discrete energy conversions afford the ocean
modeler some insurance that the numerical algorithm is performing with a degree
of physical integrity.
ρ,t + ρo ∇ · v = 0 (6.1)
z
( D ρ ),t = −ρo ∇ · U + ρw qw (6.2)
u,t + ∇ · (v uρ ) + M ẑ ∧ vρ = − f ẑ ∧ v − ∇( p/ρo ) + (ρ/ρo ) F(u) (6.3)
p,z = −ρ g (6.4)
(ρ T ),t + ρo ∇ · (v T ) = −∇ · (ρ F) + ρ S , (6.5)
where
ρ vρ = ρo v (6.6)
is the linear momentum density. For finite domains, these equations are combined
with the surface and bottom kinematic boundary conditions
ρ η,t + ρo u · ∇η = ρo w + ρw qw at z = η (6.7)
u · ∇H + w = 0 at z = − H. (6.8)
∇·v = 0 (6.9)
η,t = −∇ · U + qw (6.10)
(u)
u,t + ∇ · (v u) + M ẑ ∧ v = − f ẑ ∧ v − ∇( p/ρo ) + F (6.11)
p,z = −ρ g (6.12)
T,t + ∇ · (v T ) = −∇ · F + S . (6.13)
η,t + u · ∇η = w + qw at z = η (6.14)
u · ∇H + w = 0 at z = − H. (6.15)
where we used the relation ρ vρ = ρo v for the linear momentum density. Now take
the inner product of this budget with the horizontal velocity uρ to find
where tensor labels were exposed where needed for clarity. Use of mass conserva-
tion in the form ρ,t + ∇ · (ρ vρ ) = 0 leads to
where
1 ρ ρ
K=
u ·u (6.19)
2
is the kinetic energy per mass of a continuum model fluid parcel. Notably, it is
the model’s velocity uρ , not u, that determines the model’s kinetic energy. This
distinction is relevant only for the non-Boussinesq fluid, since for the Boussinesq
fluid u = uρ .
with hk the vertical thickness of the cell. The surface cell k = 1 has the budget
where the time derivative term is dropped for Boussinesq fluids. For surface cells
where k = 1,
ρo w z0 = −ρw qw + ρ z1 ∂t η (6.25)
w z0 = −∇ · U. (6.26)
Note that it is important not to confuse the advective velocity w z0 , diagnosed through
continuity, with the distinct vertical velocity w( z = η) used in the surface kinematic
boundary condition.
we start by noting that the projection of the horizontal velocity uρ onto the horizon-
tal pressure gradient is given by
uρ · ∇ p = vρ · ∇ p − wρ p,z
= ∇ · (vρ p) − p ∇ · vρ + ρ g wρ (6.28)
where the hydrostatic balance p,z = ρ g was used. The divergence ∇ · vρ vanishes
for a Boussinesq fluid, yet it represents a nontrivial conversion of kinetic to internal
energy for the non-Boussinesq fluid (see Griffies (2004)). Integration over the full
ocean domain yields
Z Z Z
− dV uρ · ∇ p = dA p (uρ · ∇η − wρ ) + dV ( p ∇ · vρ − wρ ρ g) (6.29)
z=η
where use was made of periodic and/or no-normal flow side boundary conditions,
as well as the bottom kinematic boundary condition. The surface kinematic bound-
ary
uρ · ∇η − wρ = ρw qw − ρ η,t (6.30)
leads to
Z Z Z
− dV uρ · ∇ p = dA p [(ρw /ρ) qw − η,t ] + dV ( p ∇ · vρ − wρ ρ g). (6.31)
z=η
88 CHAPTER 6. ENERGETICS ON THE B-GRID LATTICE
Note that mass conservation over a fluid column allows us to alternatively write
Zη
ρw qw − ρ η,t = −ρo ∇ · U + dz ρ,t (6.32)
−H
with the second term, the so-called steric contribution, absent for the volume-conserving
Boussinesq fluid (see Griffies (2004) for more discussion of steric effects). Indeed,
for the Boussinesq fluid we have (1) v = vρ , (2) ∇ · v = 0, (3) the density ρ(η)
appearing in the surface kinematic boundary condition is set to ρo , (4) η,t = −∇ ·
U + qw expressing volume conservation over a fluid column, thus leading to
Z Z Z
− dV u · ∇ p = dA p ∇ · U − dV w ρ g. (6.33)
z=η
It is useful to consider three special cases, the simplest of which is the rigid lid
Boussinesq ocean. Most rigid lid models suppress the addition of fresh water1 , in
which case ∇ · U = 0 thus rendering
Z Z
dV u · ∇ p = dV w ρ g. (6.34)
That is, for the rigid lid hydrostatic Boussinesq ocean without fresh water forc-
ing, the global effects of work done by the horizontal currents against horizontal
pressure gradients are equal to the work by vertical currents against gravity. This
equality affords the following interpretation. In a hydrostatic fluid, vertically in-
tegrated density directly determines pressure. Hence, work against a horizontal
pressure gradient force is associated with a rearrangement of the density field in
the vertical, which then involves work against gravity.
A free surface Boussinesq fluid satisfies the more general balance
Z Z Z
− dV u · ∇ p = dA p ∇ · U − dV w ρ g. (6.35)
z=η
The surface term accounts for the possibility of atmospheric pressure to apply work
to a dilatating vertical column of fluid.
For a free surface non-Boussinesq fluid, the full identity (6.37) is applicable.
The new term p ∇ · vρ accounts for the ability of pressureR
forces internal
Rη
to the
fluid to do work on dilatating fluid parcels, and the term z=η dA ( p/ρ) ( − H dz ρ,t )
accounts for work done by atmospheric pressure on a column of ocean fluid ex-
periencing expansion and/or contraction due to changes in the depth integrated
density–the so-called steric effects.
cells. With the B-grid used in MOM4, it is otherwise not obvious how to compute
these velocities. An alternative to the approach in Section 6.2.1 provides insight
for this purpose, with the distinction relevant only for the non-Boussinesq case.
Instead of uρ · ∇ p, we here consider
u · ∇ p = v · ∇ p − w p,z
= ∇ · (v p) − p ∇ · v + ρ g w. (6.36)
The divergence term ∇ · v = −ρ,t /ρo vanishes for the Boussinesq fluid. Integrat-
ing over the ocean domain and using the surface and bottom kinematic boundary
conditions leads to
Z Z Z
−ρo dV u · ∇ p = dA p (ρw qw − ρ η,t ) − ρo dV [ p (ρ,t /ρo ) + w ρ g], (6.37)
z=η
where use was made of periodic and/or no-normal flow side boundary conditions.
This result reduces to the same Boussinesq result (6.33) considered in Section 6.2.2.
The non-Boussinesq case is distinct.
+ g ∑ dau dhu [u FAY ( FAX (ρ) δi H )/dxu + v FAX ( FAY (ρ) δ j H )/dyu](6.38)
.
i, j,k
δi H = Hi+1 − Hi (6.41)
δ j H = H y+1 − H y (6.42)
90 CHAPTER 6. ENERGETICS ON THE B-GRID LATTICE
of the bottom topography. The first term in equation (6.38) is the lateral pressure
gradient taken between cells living on the same discrete k-level. The second term
arises from the use of bottom partial cells, where the depth of a k-level is generally
a function of horizontal position. This term is the same that appears in terrain-
following sigma models. In both cases, the horizontal pressure gradient is given by
two terms
∇ z p = (∇σ + ∇σ H ∂ z ) p = ∇σ p − ρ g ∇σ H, (6.43)
where the hydrostatic balance p,z = −ρ g was used to reach the second relation. The
slope of sigma surfaces can reach 1/100 next to continental slopes, at which point
the “sigma-coordinate correction term” ρ g ∇σ H can be on the order of ∇σ p. In this
case, the horizontal pressure force becomes the result of two sizable terms, each
having separate numerical errors that generally do not cancel. The result can be
spurious pressure forces that drive nontrivial unphysical currents. As described by
Pacanowski and Gnanadesikan (1998), the sigma correction term introduces only
very minor spurious currents in MOM due to (1) MOM’s use of z-levels throughout
the region above the topography, thus isolating the sigma-correction term to just the
bottom-most level, (2) The discrete horizontal pressure gradient is chosen so that
if pressure is a linear function of depth, then the discrete gradient vanishes. This
discretization choice greatly reduces the magnitude of spurious flows.
where the boundary terms drop out for either periodic or solid wall conditions, and
we introduced the backward meridional average operator
a j + a j−1
BAY ( a) = . (6.45)
2
Let us now define the zonal thickness weighted advective transport velocity on the
eastern face of a tracer cell as
where dytei, j is the meridional width of the tracer cell’s east side (see Figure 6.2 for
definitions of grid distances). Doing so leads to
where dhwti, j,k is the vertical distance between tracer points Tk and Tk+1 . The
pnk+1 w btnk boundary term vanishes since w btnk = 0; however, the surface term
p1 w bt0 is nonzero for a free surface model where w bt0 6= 0 (e.g., equations (6.53)
and (6.54)).
The discrete hydrostatic pressure is computed as an estimate of pressure at the
depth of the tracer point, rather than as an average pressure over the full tracer cell
(Figure 6.1). That is,
pk+1 = pk + g dhwtk ρk z , (6.58)
where ρk z = (ρk + ρk+1 )/2 is the vertically averaged density over the tracer cell.
Using this pressure then leads to
!
1
P1 = − ∑ dati, j p1 w bt0 + g ∑ dhwtk w btk ρk z + ∑ pk dhtk ∂t ρk . (6.59)
i, j k
ρo k
It is useful to make a correspondence with terms from the continuum given in Sec-
tion 6.2.2. First, a rigid lid Boussinesq fluid with zero fresh water flux has
!
P1 = − ∑ dati, j g ∑ dhwtk w btk ρk z , (6.60)
i, j k
where w bt0 = 0 and the density time tendency term vanishes. Hence, in this
case, work by horizontal currents on the horizontal pressure gradient equals work
against gravity. Allowing for a free surface Boussinesq fluid yields
!
P1 = − ∑ dati, j − p1 ∇ · U + g ∑ dhwtk w btk ρk z , (6.61)
i, j k
where the density time tendency term vanishes and w bt0 = −∇ · U. The added
term accounts for work done by pressure p1 at the surface tracer point on a dilatat-
ing vertical column of fluid. This pressure is given by
where eta t is the surface height on tracer cells, and dhwtk=0 is the distance from
z = 0 to the k = 1 tracer point z1 . This pressure has a contribution from the
hydrostatic pressure within the fluid layer between z = eta t and z = z1 , and that
from the overlying atmosphere. For a non-Boussinesq fluid, equation (6.53) for
w bt0 leads to
!
P1 = − ∑ dati, j ( p1 /ρo ) (ρ1 ∂t η − ρw qw ) + g ∑ dhwtk w btk ρk z + ∑ dhtk pk ∂t ρk ,
i, j k k
(6.63)
which again has a direct correspondence to the continuum results.
6.2. PRESSURE WORK CONVERSIONS 93
dhwt(k)
k+1
Figure 6.1: Schematic of the vertical grid cell arrangment used for computing the
hydrostatic pressure at a depth k + 1 in terms of the pressure at depth k using the
equation pk+1 = pk + g dhwtk ρk z . The vertical average of density is meant to ac-
count for the part of density within each of the two adjacent cells. The factor of 1/2
used in the average operator yields an approximate average when vertical cells are
non-uniform. Yet the 1/2 factor is used for all vertical grid spacing since it renders
a simple conversion of discrete pressure work to discrete gravity work.
Now consider the zonal piece of the sigma-correction term from equation (6.38)
where boundary terms vanish. Introducing the zonal thickness weighted advective
transport velocity (6.46) yields
Moving the difference operator δi H = Hi+1 − Hi from the depth H to the remaining
terms gives
P2x = − g ∑ H δi (dyte FAX (ρ) uh et) = − g ∑ H dat BDX ET ( FAX (ρ) uh et),
(6.67)
where boundary terms vanish. Similar manipulations with the meridional piece of
P2 lead to
P2 = − g ∑ H dat [ BDX ET ( FAX (ρ) uh et) + BDY NT ( FAY (ρ) vh nt)]. (6.68)
The P2 term accounts for alterations in the potential energy due to the use of partial
bottom cells.
94 CHAPTER 6. ENERGETICS ON THE B-GRID LATTICE
dxtn(i,j)
dyt(i,j)
Figure 6.2: Time independent horizontal grid distances (meters) used for the tracer
cell Ti, j in MOM4. dxti, j and dyti, j are the grid distances of the tracer cell in the
generalized zonal and meridional directions, and dati, j = dxti, j dyti, j is the area of
the cell. The grid distance dxtni, j is the zonal width of the north face of a tracer cell,
and dytei, j is the meridional width of the east face. Note that the tracer point Ti, j is
not generally at the center of the tracer cell. Distances are functions of both i and j
due to the use of generalized orthogonal coordinates.
6.2.3.5 Summary
In summary, the projection of the horizontal velocity onto the downgradient pres-
sure field is given by
+ g ∑ dau dhu [u FAY ( FAX (ρ) δi H )/dxu + v FAX ( FAY (ρ) δ j H )/dyu].
i, j,k
!
1
= − ∑ dati, j p1 w bt0 + g ∑ dhwtk w btk ρk z
+ ∑ dhtk p ∂t ρk
i, j k
ρo k
− g ∑ H dat [ BDX ET ( FAX (ρ) uh et) + BDY NT ( FAY (ρ) vh nt)] (6.69)
The MOM4 diagnostic energy conversion error has proven to be quite useful for de-
tecting improper discretization of various algorithms. That diagnostic computes
the left hand side of equation (6.69) and compares to the right hand side. Differ-
ences are due to errors in the code. The reason this diagnostic is so effective is that
it involves advective velocities on the tracer cells, both tracer and velocity cell dis-
tances, the calculation of pressure, and details of partial cells. Each require precise
discretization in order to ensure an energy conversion error at the roundoff level.
6.3. KINETIC ENERGY ADVECTION 95
The advection metric term drops out trivially. For clarity, we selectively expose
tensor labels on the horizontal velocity components to rewrite the divergence term
as
where the continuity equation ρ,t = −ρo ∇ · v was used in the last step. For a
Boussinesq fluid, similar manipulations lead to
um ∇ · (v um ) = ∇ · (v K), (6.71)
For a rigid lid Boussinesq model with zero fresh water forcing, A B = 0. More
general models have nonzero A due either to boundary contributions and/or di-
latations of fluid parcels/columns.
In this equation,
dh1dy = ∂ y ln dx (6.75)
dh2dx = ∂ x ln dy (6.76)
are the model arrays carrying information about the partial derivatives of the grid
spacing in the two orthogonal directions. The sum in equation (6.74) vanishes triv-
ially at each grid point upon writing out the two terms.
We next consider the scalar product of the horizontal convergence term with the
horizontal velocity uρ , and integrate over the full ocean
− Ahorz =
∑ dau dhu u ρ ρ
· BDX EU (uh eu ∗ FAX (u )) + BDY NU (vh nu ∗ FAY (u )) /dhu.ρ
i, j,k
(6.77)
These operators are the unweighted averages used to estimate velocity on the ve-
locity cell faces. They are used to define the centered difference advective fluxes of
velocity. MOM4 also uses the backward derivative operators
These backward derivative operators act on fields defined at the east and north face
of velocity cells, respectively (see Figure 6.3 for definitions of grid distances).
As detailed in Chapter 5, thickness weighted horizontal advective velocities
uh eu and vh nu are defined in MOM4 by remapping the horizontal advective ve-
locities uh et and vh nt, defined in Section 6.2.3.1, onto the velocity cell faces. They
6.3. KINETIC ENERGY ADVECTION 97
∑ uρi · dyuei uh eui uρi+1 + dyuei uh eui uρi − dyuei−1 uh eui−1 uρi − dyuei−1 uh eui−1 uρi−1
It is appropriate to set
u Nk+1 = 0 (6.87)
since k = Nk + 1 is interpreted as part of the solid earth, thus leading to
Nk
−2 Ainterior
vert = ∑ dau w bu1 (uρ2 · uρ1 ) + 2 ∑ dau Kk (w buk−1 − w buk ).
k=2
For the surface cell, reference to the semi-discrete momentum budget (6.21) leads
to
k=1
−2 Avert = − ∑ dau w bu1 uρ1 · (uρ1 + uρ2 ) − ρ−
o
1
∑ dau uρ1 · (−uρz1
ρ1 η,t + ρw qw uρw )
(6.88)
To bring this expression more in line with the continuum results of Section 6.3.1, let
us write
uρw = (uρw − uρ1 ) + uρ1 (6.90)
to yield
Nk
Avert = ∑ dau Kk (w buk − w buk−1 ) + ∑ dau K1 w bu1
k=2
+ ρ−
o
1
∑ dau K1 (−ρ1 η,t + ρw qw ) + (2 ρo )−1 ∑ dau uρ1 · (uρw − uρ1 ) ρw qw
Nk
= ∑ dau Kk (w buk − w buk−1 ) + (2 ρo )−1 ∑ dau uρ1 · (uρw − uρ1 ) ρw qw
k=1
(6.91)
where the continuity equation (6.82) was used to reach the second equality. For the
Boussinesq model with zero fresh water forcing, or with uρw − uρ1 = 0, then Ahorz +
Avert = 0, whereas the more general case is nonzero due to the compressibility of
seawater and/or fresh water forcing.
6.4. KINETIC ENERGY IN THE EXTERNAL AND INTERNAL MODES 99
dxun(i,j)
dyu(i,j)
U(i,j)
dyue(i,j)
dxu(i,j)
Figure 6.3: Time independent horizontal grid distances (meters) used for the ve-
locity cell Ui, j in MOM4. dxui, j and dyui, j are the grid distances of the velocity cell
in the generalized zonal and meridional directions, and daui, j = dxui, j dyui, j is the
area of the cell. The grid distance dxuni, j is the zonal width of the north face of
a velocity cell, and dyuei, j is the meridional width of the east face. Note that the
velocity point Ui, j is not generally at the center of the velocity cell. Distances are
functions of both i and j due to the use of generalized orthogonal coordinates.
b = ψ − ψz .
ψ (6.94)
Using this notation, the horizontal velocity is split into its external and internal
modes
u = uz + ub. (6.95)
Substituting these velocities into the kinetic energy per unit mass yields
1 z z
K= (u · u + u b + 2 uz · u
b·u b ). (6.96)
2
The depth averaged kinetic energy per mass is given by
z 1 z z z
K = (u · u + u
b·u
b ). (6.97)
2
100 CHAPTER 6. ENERGETICS ON THE B-GRID LATTICE
Note the decoupling of the external and internal modes when vertically integrating.
Hence, the depth averaged kinetic energy per mass can be thought of as the sum of
a contribution from the external mode kinetic energy per mass
1 z z
Kext = u ·u , (6.98)
2
and the depth averaged internal mode kinetic energy per mass
z 1 z
Kint = b·u
u b . (6.99)
2
In particular, this means that when computing the terms contributing to the evolu-
tion of global kinetic energy, it is sufficient to take the scalar product of the terms in
the equation of motion with the internal and external mode velocities, respectively,
and then volume average. Notably, surface pressure gradients contribute only to
the external mode energy.
Contents
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
7.2 Continuum model budget . . . . . . . . . . . . . . . . . . . . . . . 101
7.3 Discrete Boussinesq rigid lid budget . . . . . . . . . . . . . . . . 102
7.4 Discrete non-Boussinesq free surface budget . . . . . . . . . . . 104
7.4.1 Separate time stepping for tracer, density, and thickness . 105
7.4.2 Concerning the tracer and baroclinic time steps . . . . . . 106
7.4.3 MOM4 diagnostics for tracer mass . . . . . . . . . . . . . . 106
7.5 Comments on three-time level schemes . . . . . . . . . . . . . . . 108
The purpose of this chapter is to interpret the evolution of total ocean tracer con-
tent. It is noted that a fundamental problem of time filtered leap-frog time stepping
schemes is the inability to exactly conserve total tracer content.
7.1 Introduction
Time evolution of the total ocean tracer is determined by its input through the
ocean boundaries and source/sink terms within the interior. Upon time discretiza-
tion, the definition of total ocean tracer becomes dependent on details of the time
stepping scheme. Furthermore, the introduction of explicit time filters, such as the
Robert-Asselin filter used with the leap-frog scheme or time averaging used in the
MOM4 explicit free surface method, alters the evolution of total tracer in a non-
conservative manner. The goal of this section is to precisely define what we mean
by total ocean tracer content in light of time discretization details.
Note that for a Bouissinesq fluid, tracer concentration remains a scalar if it is inter-
preted as the tracer mass per volume of a water parcel, since volume is a scalar for
Boussinesq fluids.
R
In this case, the total tracer mass for a Bouissinesq fluid is given
by MT = ρo dV T.
Time evolution of the total tracer mass is determined by sources and sinks of
tracer within the fluid and at the oceanic boundaries. Using the mass and tracer
budgets given in Chapter 3 leads to
Z
∂ t MT = ∂ t (ρ dV ) T
Z Zη
= dA (ρ T ) z=η η,t + dz (ρ T ),t
−H
Z Z
= dA T (ρ η,t − ρo N
b · v) − N
b · ρ F + (ρ dV ) S
z=η
Z Z (7.2)
= dA T ρw qw − N
b · ρ F + (ρ dV ) S
z=η
Z Z
= dA T ρw qw − N · ρ F + (ρ dV ) S
b
z=η
Z Z
=− dA ρ Q T + (ρ dV ) S ,
z=η
where we used the surface kinematic boundary condition, assumed zero flux through
the solid boundaries and/or periodic lateral boundary conditions, and set the sur-
face tracer flux equal to that coming in from boundary layer model.
where w z0 = 0 for the rigid lid. Integrating over the total model volume cancels the
horizontal fluxes so long as there are no tracer sources, such as geothermal heating,
at the solid boundaries. The volume integrated vertical flux convergence reduces to
an area integral of the surface tracer flux. Hence, exposing some model grid fields
and discrete indices leads to the time tendency
!
∂t ∑ dat dht T = − ∑ dat QT = − ∑ dat Qturb
T (7.4)
i, j,k i, j i, j
7.3. DISCRETE BOUSSINESQ RIGID LID BUDGET 103
where dhti, j,k is the tracer cell thickness, dati, j = dxti, j dyti, j is the horizontal area
of a tracer cell, and indices are exposed only when needed. Multiplying both sides
by the constant density ρo renders the budget for total tracer mass in the rigid lid
Boussinesq fluid.
The balance (7.4) takes on various forms depending on how time is discretized.
Following the approach in MOM, we introduce a leap-frog time stepping scheme
for the tendency of tracer concentration
where
with positive F indicating tracer added to the ocean domain. T represents the total
mass of tracer if T represents mass per volume, or the total ocean heat divided by
C p if T is the potential temperature θ. The budget for total tracer mass now takes
the more tidy form
T (τ + ∆τ ) = T R (τ − ∆τ ) + 2 ∆τ F (τ ). (7.9)
Because of the Robert time filter, tracer mass at τ + ∆τ does not equal that at τ − ∆τ
plus that input through the ocean surface. That is, time filtering compromises the
discrete balance of total ocean tracer mass. It is as if the flux input via the surface
term F (τ ) were modified, with the magnitude of the modification depending on
the amount of time filtering applied. This is not a satisfying situation. Yet it is a
necessary consequence of using the time filtered leap-frog scheme. Such will also
be the case with the free surface, where there is a time filter applied to both the
tracer and surface height, thus further compromising the integrity of the conserva-
tion properties. Two-time level schemes, such as predictor-corrector, avoid these
problems via the use of implicit time dissipation, and so are preferable from the
present perspective. Yet we will have more to say on two-time level schemes in
Section 7.5.
104 CHAPTER 7. TOTAL OCEAN TRACER CONTENT
To pursue this point a bit further, consider the special case of zero surface tracer
flux for all time, for which case
T (τ + ∆τ ) = T R (τ − ∆τ ). (7.10)
Hence, there is an apparent time evolution of total tracer content even when there
are no surface fluxes. However, upon assuming T constant for all earlier time steps,
one can show inductively that T remains constant for all future time steps. As
shown in the next section, the nonlinear product of the time dependent surface
thickness and time dependent tracer concentration, in the presence of separately
applied time filters, leads this conservation statement to be compromised with a
free surface.
where the area integrated surface tracer flux is given by the sum of the turbulent
and fresh water terms
The density weighting, fresh water term, and time dependent surface cell thickness
each provide added features relative to the rigid lid Boussinesq case.
Performing a leap-frog time step on the thickness weighted tracer concentration
leads to the conservation statement
T (τ + ∆τ ) = T (τ − ∆τ ) + 2∆τ F (τ ), (7.13)
where
T (τ ) ≡ ∑ dat dht(τ ) ρ(τ ) T (τ ) (7.14)
i, j,k
is the domain integrated tracer mass, and time filtering is briefly ignored. This
conservation statement has the same form as that for the rigid lid. In both cases,
the use of time filters compromises our familiar interpretation of the discrete time
budget for total tracer. Additionally, because thickness must be time stepped sep-
arately from the thickness weighted tracer, time filtering is also applied separately.
Hence, upon introducing time filtering, the nonlinear product of thickness times
tracer concentration further compromises the integrity of the conservation state-
ment. No such issue arises with density, and hence total ocean mass, since its time
stepping is diagnosed by the extrapolation method described in Greatbatch et al.
(2001) and Griffies (2004).
7.4. DISCRETE NON-BOUSSINESQ FREE SURFACE BUDGET 105
This time discretization has the same accuracy as when discretizing the product
h ρ T as a single object. However, the conservation statement is a bit more compli-
cated in the present approach, as seen in the following. Note that dht(τ + ∆τ ) −
dht(τ − ∆τ ) is nonzero only for the surface grid cell. However, for purposes of
symmetry, it is useful to keep it around for all depths.
To proceed, rearrange the previous result and indicate where the time filtered
fields are used in the model’s time stepping.
∑ dat [dht(τ ) ρ(τ + ∆τ ) T (τ + ∆τ ) + dht(τ + ∆τ ) ρ(τ ) T (τ )]
i, j,k
− [dht F (τ − ∆τ ) ρ(τ ) T (τ ) + dht(τ ) ρ(τ − ∆τ ) T R (τ − ∆τ )] = 2 ∆τ F (τ ),
(7.16)
where there is no time filtering applied to the density field since it is updated via an
extrapolation method. Furthermore, as detailed in Griffies (2004), the time filtered
ocean surface height takes the form of a time average over the barotropic cycle
N
1
η F (τ ) = η(b) (τ ) = ∑
N + 1 n=0
η(b) (τ − ∆τ , tn ). (7.17)
To isolate the effects of time filtering on the tracer content, we find it convenient
to write
= 2 ∆τ F (τ ). (7.18)
To bring this equation into a more tidy form, define the two-time level tracer mass
2 T (τ , τ + ∆τ ) =
∑ dat [dht(τ ) ρ(τ + ∆τ ) T (τ + ∆τ ) + dht(τ + ∆τ ) ρ(τ ) T (τ )] (7.19)
i, j,k
106 CHAPTER 7. TOTAL OCEAN TRACER CONTENT
2 T F (τ , τ + ∆τ ) =
∑ dat [dhtF (τ ) ρ(τ + ∆τ ) T (τ + ∆τ ) + dht(τ + ∆τ ) ρ(τ ) T R (τ )], (7.20)
i, j,k
T (τ , τ + ∆τ ) = T (τ − ∆τ , τ ) + ∆τ F (τ ) + [T F (τ − ∆τ , τ ) − T (τ − ∆τ , τ )]. (7.21)
step, this diagnostic reveals the extent of the associated tracer non-conservation.
Such is part of the tracer-change diagnostic in MOM4.
In addition to checking the left and right hand sides of equation (7.16) over a
single time step, it is important to check how well the ocean inputs the tracer fluxed
through its surface over a period of time. That is, does the tracer content at some
given time equal to that at an earlier time plus the total tracer input through the
surface during the intermediate times? We already answered this question in the
negative for the rigid lid, in which case the mis-match arises from the use of time
filters. The negative answer applies also for the free surface, again because of time
filters as seen by equation (7.21). We now consider details in order to formulate the
MOM4 diagnostic which determines the magnitude of the mis-match.
For this purpose, integrate equation (7.21) over a finite time N ∆τ to obtain
N
T (τ + N∆τ , τ + ( N + 1)∆τ ) = T (τ , τ + ∆τ ) + ∆τ ∑ F (τ + n∆τ )
n=1
N
+ ∑ [T F (τ + (n − 1) ∆τ , τ + n ∆τ ) − T (τ + (n − 1) ∆τ , τ + n ∆τ )]. (7.22)
n=1
where A = ∑ dat is the total tracer cell area at the ocean surface. Second, compute
an effective flux
1
error(2) = ×
N ∆τ A
N
[T (τ + N∆τ, τ + ( N + 1)∆τ ) − T (τ , τ + ∆τ ) − ∆τ ∑ F (τ + n∆τ )] (7.24)
n=1
∆τ ∆τ ∆τ ∆τ ∆τ ∆τ
Figure 7.1: Schematic of time stepping for tracer concentration and total tracer mass
in the case where N = 6. The lower branch shows the N over-lapping 2∆τ leap-
frog steps used for updating the tracer field. The upper branch shows the N time
steps of length ∆τ moving the total tracer mass over the half-integer τ levels. Note
that the increments for τ are in units of ∆τ.
suggestions.
The two-time level predictor-corrector schemes contain time damping applied
in a semi-implicit manner. Hence, there is no need to employ a Robert filter. Careful
discretization in the C-grid Hallberg Isopycnal Model Hallberg (1997) allows for an
exact conservation of total tracer if the predictor-corrector is applied to the tracer
fields as well as within the barotropic sub-cycle. The surface height that is used
to maintain conservation is driven by the barotropically time averaged vertically
integrated transport.
In an attempt to see whether a two-time level scheme is warranted for MOM,
we implemented a predictor-corrector approach for the barotropic system (similar
to Hallberg’s), yet maintained the leap-frog for the tracer and baroclinic velocity.
Unfortunately, the model was very unstable, likely due to the B-grid computational
mode, unless a time averaged surface height was incorporated at some stage in the
barotropic algorithm (see Griffies et al. (2001) for comments on such instabilities).
Yet the use of time averaging no longer allows for exact tracer conservation, even if
the tracer equation is changed to a predictor-corrector. Hence, we have found little
motivation in modifying the leap-frog scheme in the B-grid MOM. Conversion to
a C-grid may lead to more positive results, so long as the Coriolis null mode does
not then become pernicious (Adcroft et al. (1999)).
Even with the rigid lid, a leap-frog compromises the total tracer conservation.
Short of a fully two-time level scheme, it has been common to remove the leap-frog
splitting mode by periodically applying a two-time level Euler step instead of us-
ing a time filter. Extensive experience with this approach has prompted modelers to
abandon the use of such special-time step “tricks”, largely due to subtleties related
to vertical adjustment processes (see Griffies et al. (2000a) for a review and refer-
ences). A predictor-corrector for a rigid lid B-grid model thus makes more sense,
because the issues of stabilizing the free surface are no longer relevant and all time
steps are treated the same. Because the rigid lid is not suitable for other reasons
(e.g., Griffies (2004)), there has been no attempt to see how useful a two-time level
B-grid MOM would be.
110 CHAPTER 7. TOTAL OCEAN TRACER CONTENT
Part III
Shortwave heating
Contents
where I0− , in units of W m−2 , is the total shortwave downwelling radiative heating
per unit area incident at the earth surface, and F ( z) is a dimensionless attenuation
function. Note that the total downwelling radiation I0− is to be distinguished from
the total shortwave heating I0 , where I0− = (1 − α ) I0 , with α ≈ 0.06 the sea surface
albedo.
Shortwave heating affects the heat budget locally according to
In this equation, F z accounts for vertical processes such as advection and diffusion,
c p is the heat capacity of seawater, ρ is the in-situ density which for a Boussinesq
fluid is set to the Boussinesq reference density ρo .
Shortwave heating leads to the following net heat flux over a column of ocean
fluid
Zη
(ρo /c p ) dz ∂ z I = (ρo /c p ) [ I (η) − I (− H )]. (8.3)
−H
where I (η) is often approximated as I (0). We assume there is no shortwave heating
of the solid rock underneath the ocean fluid, so I ( z = − H ) = 0 is appropriate,
with this boundary condition set via masks in MOM4. Although the expression
(8.5) suggests the upper boundary condition I (0) = I0− , we must be careful to not
double-count the shortwave source in MOM4 since it is typically also carried as
part of the surface temperature flux array st f . We now present the two approaches
available in MOM4.
The vertical convergence of penetrative shortwave radiation, (ρo /c p )∂ z I, is in-
corporated into MOM4’s potential temperature equation via a source term. Addi-
tionally, it is typical to include the total downwelling shortwave heating I0− within
the surface flux array st f , where other forms of heating such as those from latent
and long-wave affects are also incorporated. Hence, for proper accounting of the
shortwave heating, the upper boundary condition for the irradiance function must
be specified as
0 if I0− is already included in st f
I (η) = (8.4)
I0− ( x, y) if I0− is NOT already included in st f .
The typical practice at GFDL is to set I (η) = 0 since I0− is already included in st f .
Care should be exercised by those using the opposite convention.
model. Thus, the resulting equation for the infared portion of the downwelling
radiation is:
I IR ( x, y, z) = I IR− ( x, y) e−z/(0.267 cos θ) (8.6)
where again θ = 0. This relationship assumes z is the depth in meters and that
I IR− ( x, y) is a fraction, of the total downwelling radiation, I0− such that
and
FIR + FV IS = 1 (8.8)
where FIR and FV IS are the fractions of infared (750nm to 2500nm) and visible
(300nm to 750nm) radiation downwelling from the surface ocean. Although Morel
and Antoine (1994) note that water vapor, zenith angle, and aerosol content each
can effect the fraction of incoming radiation that is represented by infared and vis-
ible light, in the present implementation we have chosen to keep these fractions
constant such that FIR = 0.46.
The second and third exponentials represent a parameterization of the attenua-
tion function for downwelling radiation in the visible range (300nm -750nm) in the
following form:
This form further partitions the visible radiation into long (V1 ) and short (V2 ) wave-
lengths assuming
V1 + V2 ≡ 1. (8.10)
V1 , V2 , ζ1 and ζ2 are calculated from an empirical relationship as a function of
chlorophyll-a concentration using methods from Morel and Antoine (1994). Through-
out most of the ocean V1 < 0.5 and V2 > 0.5. The e-folding length scales ζ1
and ζ2 are the e-folding depths of the long (ζ1 ) and short visible and ultra vio-
let (ζ2 ) wavelengths. Based on the chlorophyll-a climatology used in the GFDL
models, ζ1 should not exceed 3m while ζ2 will vary between 30m in oligotrophic
waters and 4m in coastal regions. All of these constants are based on satellite es-
timates of chlorophyll-a plus Pheaophytin-a, as well as parameterizations which
have “nonuniform pigment profiles” (Morel and Antoine (1994)). The ”nonuniform
pigment profiles” have been proposed to account for deep chlorophyll maxima
that are often observed in highly stratified oligotrophic waters (Morel and Berthon
(1989)).
Contents
The purpose of this chapter is to present the vertical adjustment schemes avail-
able in MOM4.
9.1 Introduction
The hydrostatic approximation necessitates the use of a parameterization of vertical
overturning processes. The original parameterization used by Bryan in the 1960’s
was motivated largely from ideas then used for modeling convection in stars (Bryan
(1969)). Work by Marshall and collaborators (Klinger et al. (1996), Marshall et al.
120 CHAPTER 9. VERTICAL ADJUSTMENT SCHEMES
(1997)) have largely indicated that the basic ideas of vertical adjustment are useful
for purposes of large-scale ocean circulation.
The Cox (1984) implementation of convective adjustment (the “NCON” scheme)
may leave columns unstable after completing the code’s adjustment loop. Various
full convective schemes have come on-line, with that from Rahmstorf (1993) imple-
mented in MOM4. An alternative to the traditional form of convective adjustment
is to increase the vertical mixing coefficient to some large value (say ≥ 10m2 s−1 ) in
order to quickly diffuse vertically unstable water columns. Indeed, it is this form
recommended from the study of Klinger et al. (1996), and it is the approach com-
monly used in mixed layer schemes such as Pacanowski and Philander (1981) and
Large et al. (1994).
• Implicit vertical mixing: By setting the namelist parameter aidi f = 1.0, all
vertical diffusion is handled implicitly. There are two approaches depending
on the vertical mixing scheme used.
1. When using the constant vertical mixing module, the vertical diffusiv-
ity is set to a maximum value determined by a namelist di f f cbt limit
upon reaching a gravitationally unstable situation. di f f cbt limit = 10.0
(MKS) is a typical value.
2. When using the Pacanowski and Philander or KPP vertical mixing scheme,
both the vertical diffusivity and vertical viscosity are set to the namelist
settings di f f cbt limit and visc cbu limit upon reaching a gravitationally
unstable situation. di f f cbt limit = visc cbu limit = 10.0 (MKS) are typ-
ical values.
kappa m. If wishing to adjust via large vertical diffusivities, then set di f f cbt limit
to a large value as described above, and set the namelist convective ad just = . f alse.
ρo u = ρ uρ . (9.1)
Hence, for the non-Boussinesq model, it is uρ that is mixed implicitly in time when
using the implicit mixing scheme. There is no distinction between uρ and u for the
Boussinesq model.
where
F z = −κ T,z (9.3)
is the downgradient vertical tracer flux. In MOM4, we prefer to separate the density
and tracer concentration prior to temporal discretization, thus to isolate the time
tendency of the tracer concentration
The density time tendency term vanishes for the Boussinesq case, and the factor
ρo /ρ is set to unity. For the non-Boussinesq case, the term −( T /ρ) ρ,t provides an
added forcing to the tracer concentration, and the ρo /ρ factor modifies the vertical
diffusion operator. The modification to the vertical diffusion operator is considered
in the following sections where we discuss the implicit treatment of this equation
via either implicit vertical diffusion or convective adjustment.
aidi f 2 ∆τ
Γk (τ ) = (9.6)
(ρk (τ )/ρo ) dhtk (τ )
which appears quite frequently in the following discussion. The profile φ∗k (τ + 1)
represents the tracer at the updated model time τ + 1 arising from all processes
treated explicitly in time. The leap-frog time step 2 ∆τ is generally larger than that
allowed by CFL stability given the value of the vertical diffusivity, hence the need to
solve the equation implicitly in time. The implicit vertical mixing parameter aidi f
is set to unity for the fully implicit method. The field dhtk is the vertical thickness
of the tracer cells at time τ. For vertical mixing of velocity, dht becomes the velocity
cell thickness dhu.
The influence of the non-Boussinesq formulation can be thought of as modi-
fying the vertical thickness dhtk of the tracer cells to (ρk (τ )/ρo ) dhtk (τ ) (e.g., see
equation (9.6)). Notably, this modification holds whether we are considering ver-
tical diffusion or vertical convective adjustment, since it is independent of vertical
diffusivity.1
φk (τ + 1) − φk+1 (τ + 1)
Fkz = −κk . (9.7)
dhwtk
1 For these purposes, vertical convective adjustment can be thought of as vertical diffusion with an
infinite diffusivity.
9.5. IMPLICIT VERTICAL MIXING 123
The factors of φ are evaluated at time τ + 1 because of the implicit treatement. The
vertical mixing coefficient κk has a general space-time dependence set by a vertical
mixing scheme. As for the flux itself, the diffusivity κk is situated at the bottom of
the tracer or velocity cell, depending on whether φ is a tracer field or velocity com-
ponent. The array dhwtk represents the vertical distance between tracer points at
time τ. For vertical mixing of velocity, dhwt becomes the distance between velocity
points dhwu.
At the ocean surface, the vertical flux is given by the surface boundary condition
s f lux placed on the velocity or tracer. For a tracer,
with st f MOM4’s surface tracer flux array with units of velocity times tracer con-
centration. The minus sign arises from the MOM4 convention that associates a
positive st f with an increase in tracer within the k = 1 cell. In contrast, the present
discussion assumes a convention for the flux F z whereby a positive Fkz=0 is associ-
ated with a decrease in tracer within the k = 1 cell. For velocity,
with sm f the surface momentum flux with units of squared velocity. At the ocean
bottom, a similar condition leads to
which leads to
Γ (τ ) κk Γk (τ ) κk
φ∗k (τ + 1) + Γk (τ ) st f = φk (τ + 1) 1+ k − φk+1 (τ + 1) .
dhwtk dhwtk
(9.13)
For velocity mixing, st f becomes sm f , and dhwt becomes dhwu.
renders
Φ∗k = Ak φk−1 (τ + 1) + Bk φk (τ + 1) + Ck φk+1 (τ + 1). (9.22)
Vertical advection
Contents
Vertical advection has been found to constrain the model time step for certain
global model configurations moreso than other of transport processes. The purpose
of this chapter is to present two methods to relax the vertical CFL constraint. The
first method time steps second order centered advection implicitly in time. The
second method is based on ideas presented by Lin and Rood (1996).
MATERIAL IN THIS CHAPTER HAS YET TO BE IMPLEMENTED IN MOM4
where we focus on tracer concentration for the moment, with trivial generalization
available for treating vertical velocity transport. For diffusive transport, as treated
in Chapter 9,
F z = −κ T,z (10.2)
is the downgradient vertical tracer flux. For advection, as treated in this chapter,
Fz = w T (10.3)
128 CHAPTER 10. VERTICAL ADVECTION
is the vertical advective flux. In MOM4, we prefer to separate the density and tracer
concentration prior to temporal discretization, thus to isolate the time tendency of
the tracer concentration
T,t = −( T /ρ) ρ,t − (ρo /ρ) ∂ z F z . (10.4)
The density time tendency term vanishes for the Boussinesq case, and the factor
ρo /ρ is set to unity. For the non-Boussinesq case, the term −( T /ρ) ρ,t provides an
added forcing to the tracer concentration, and the ρo /ρ factor modifies the vertical
transport operator. The modification to the vertical transport operator is considered
in the following sections where we discuss the implicit treatment of this equation.
Tk (τ + ∆τ ) = Tk∗ (τ + ∆τ )
( adv)
+ (Γk /2) [w btk ( Tk (τ + ∆τ ) + Tk+1 (τ + ∆τ )) − w btk−1 ( Tk−1 (τ + ∆τ ) + Tk (τ + ∆τ ))] .
(10.9)
130 CHAPTER 10. VERTICAL ADVECTION
We can put these results into a standard form via manipulations similar to those
used for vertical diffusion in Section 9.5.3. For this purpose, some rearrangement
brings equation (10.7) into the form
( adv) ( adv)
Tk∗ (τ + ∆τ ) = Tk (τ + ∆τ ) 1 − (Γk /2) w k
bt + ( Γ k / 2 ) w k−1
bt
( adv) ( adv)
+ Tk−1 (τ + ∆τ ) (Γk /2) w btk−1 − Tk+1 (τ + ∆τ ) (Γk /2) w btk . (10.10)
This equation holds for all grid points in the vertical, starting from k = 1. However,
it is useful to distinguish the treatment at the top and bottom boundaries. For k = 1,
the w btk=0 contribution is handled via a separate portion of the update process
associated with surface height time tendencies. So this term does not contribute
here, thus leading to
Tk=1 (τ + ∆τ ) = Tk∗=1 (τ + ∆τ )
( adv)
+ (Γk=1 /2) [w btk=1 ( Tk=1 (τ + ∆τ ) + Tk=2 (τ + ∆τ )) − 0], (10.11)
A similar equation results for vertical advection of velocity at the k = 1 velocity cell.
At the bottom of a tracer column, w btk=kmt generally vanishes due to volume/mass
conservation, thus leading to
Tk=kmt (τ + ∆τ ) = Tk∗=kmt (τ + ∆τ )
( adv)
+ (Γk=kmt /2) [0 − w btk=kmt−1 ( Tk=kmt−1 (τ + ∆τ ) + Tk=kmt (τ + ∆τ ))], (10.13)
At the bottom of a column of velocity points, w buk=kmu generally does not vanish,
since this velocity does not sit at the ocean bottom when there is nontrivial topo-
graphic slope. Hence, both advective fluxes are present for the velocity equation at
k = kmu. The issue of vertical velocities at the ocean bottom is discussed in Section
22.3.3 of Pacanowski and Griffies (1999).
Following the diffusion discussion in Section 9.5.3, we introduce
(
( adv)
2 Ak =
Γk (τ ) w btk−1 if k > 1 (10.15)
0 if k = 1
Bk = 1 + Ak + Ck (10.16)
( adv)
2 Ck = −Γk (τ ) w btk (10.17)
10.4. IMPLICIT FIRST ORDER UPWIND ADVECTION 131
to render
Tk∗ (τ + ∆τ ) = Ak Tk−1 (τ + ∆τ ) + Bk Tk (τ + ∆τ ) + Ck Tk+1 (τ + ∆τ ). (10.18)
Note that higher order advection schemes lead to larger stencils. So, if time stepped
implicitly, these schemes require a more expensive matrix inversion.
k−1
(z)
F(k−1)
(z)
F(k)
k+1
Figure 10.1: Schematic illustrating the vertical advective transport passing across
the faces of a model grid cell.
Part IV
Contents
11.1 Introduction
There are regions where ice-sheets have yet to form, yet ice is present in the liq-
uid water. This slurry of water-ice is commonly known as frazil. The liquid water
portion of frazil has a temperature given by the freezing temperature, which is a
function of the salinity and pressure.
During the formation of frazil, latent heat and salt is released to the liquid sea-
water. To get a rational representation of the salt added to the ocean model requires
that it be coupled to a sea ice model. For the present purposes, we describe a sim-
ple means to represent the heating effect. Note that when implementing a “frazil”
138 CHAPTER 11. CONSIDERATIONS FOR ICE-OCEAN MODELING
option in MOM4 without sea ice, we are led to the unphysical circumstance of hav-
ing sea ice formation stabilize the ocean water column. Hence, one should be quite
careful not to over interpret solutions where frazil heating occurs without the cor-
responding salinification.
Water with higher salinity freezes at lower temperatures than water with lower
salinity. If the model’s prognostic temperature equation (i.e., advection, diffusion,
surface cooling) predict that T (◦ C) < −0.054 s(psu), two effects occur. First, heat
is input to the liquid ocean water as it is extracted from the freezing ocean water.
The amount of heat given to the liquid is
where
dV = dx dy (∆z + η) (11.3)
is the surface grid cell volume, ρ is the in situ ocean density, and C p is the heat
capacity of seawater with
ρo C p = 1035 kg m−3 × 3985 Joule ◦ C−1 kg−1 = 4.125 × 106 Joule m−3 C−1 . (11.4)
As defined, Q f razil is the heat sufficient to raise the sea surface temperature to
T (◦ C) = −0.054 s(psu). Note that as the frazil heat is given to the liquid ocean
over the course of a tracer leap-frog time step, the rate of heating of the liquid due
to frazil formation is given by
with T ∗ (τ + dtts) the potential temperature updated to the new time step due to
the effects of processes other than frazil formation.
We determine the frazil heat Q f razil (Joule) within the ocean model. This heat
is applied to the ocean over a single leap-frog time step of length 2 dtts. For heat
conservation within the coupled ocean and sea ice system, this same amount of
heat must be extracted from the sea ice model. Importantly, if the sea ice model
uses a different time stepping scheme than MOM4’s leap-frog, then the frazil heat
Q f razil extracted from the sea ice model must be adjusted accordingly. For example,
the GFDL sea ice model (Sea Ice Simulator) uses a single-time step scheme rather
than a leap-frog. So the frazil heat Q f razil extracted from the sea ice model over a
11.3. ICE WITH A FREE SURFACE IN A Z-MODEL 139
single dtts time step within the ice model will be given by one-half that input to the
ocean model over a single ocean leap-frog time step 2 dtts
| Q MOM4 SIS
f razil (Joule)| = 2 | Q f razil (Joule)| . (11.7)
An example helps to illustrate the magnitude of this heating. Consider the cool-
ing of a surface grid cell with salinity 35psu, and let the cooling be such that the
model’s prognostic temperature equation predicts an updated surface cell temper-
ature of −2 ◦ C. Since −2 < −0.054 × 35 = 1.89, frazil will form. The latent heat
given to the liquid ocean associated with the frazil formation is
Q f razil (Joule) = |1.89 − 2| dV ρo C p , (11.8)
where we made the Boussinesq approximation and so set the in situ density to the
reference density ρo . For a surface grid cell of volume 105 m × 105 m × 10 m, the
formation of frazil leads to a local heat given to the ocean of
Q f razil = 4.5 × 1016 Joule. (11.9)
If this heating occurs over the course of a day (i.e., 2 dtts = 1 day), then the rate of
heating is given by
dQ f razil /dt = 4.25 × 1011 Watt. (11.10)
If the surface area dA of the grid cell is 105 m × 105 m, this heating by frazil con-
tributes
(dQ f razil /dt)/dA = 42.5 Watt m−2 . (11.11)
qw
z= η
D=H+ η
U U
Figure 11.1: The balance of volume within a column of Boussinesq fluid extending
from the free surface to the ocean bottom is represented by η,t = −∇ · U + qw ,
where η is the time dependent deviation of the surface height from its resting state
at z = 0. Time dependent fluctuations (η,t ) are driven by (1) the convergence of
volume fluxes associated with the vertically integrated horizontal currents, and (2)
the flux of volume across the ocean surface from fresh water.
is the initial thickness of ocean liquid in the grid cell. For the purposes of deriving
the ice-ocean mass balance, we are not concerned with the in situ density of ocean
fluid, but the in situ density of fresh water in the surface grid cell. That is, we are
concerned with the mass of fresh water in the cell
As determined by the dynamical ice model, some fraction of this fresh water freezes
into sea ice with density ρice , thus forming ice of mass
where hice is the ice thickness. Assuming all the fresh water mass used to form ice
comes locally from an ocean model grid cell, ice formation reduces the thickness of
liquid ocean in this grid cell. At equilibrium, the thickness of ocean fluid is therefore
given by
(∆z + η f inal ) = (∆z + ηinitial ) − hice (ρice /ρ f resh ). (11.17)
z= η
z= η
h(ocean)
h(ocean)
z=z1 z=z1
Figure 11.2: Freezing of liquid ocean to form sea ice. Shown here is a single surface
model grid cell with vertical thickness ∆z + η, where ∆z = | z1 |. The left panel
shows a cell filled with liquid ocean of some thickness hocean . The right panel shows
a cell with ice floating on liquid ocean in an equilibrium state after the transfer of
fresh water from ocean to ice has occured.
For the volume budget, the ice model provides the ocean model with a rate at
which fresh water is transferred between the sea ice and ocean. That is, the ice
model determines the ice contribution to the fresh water flux qw to be applied to
the surface height equation (11.12).
For the momentum budget, we consider the pressure applied at z = 0. This
pressure drives the vertically integrated momentum field in the model. Without
ice, Figure 11.2 shows that p z=0 is set by the atmospheric pressure as well as the
hydrostatic pressure arising from the layer of fluid between z = 0 and z = η
where again ρocean is the in situ ocean fluid density. In the presence of ice, pressure at
z = 0 has a contribution due to the mass of ice floating on the ocean. This pressure
is given by the weight per area of ice, thus leading to
The overall dynamical effect of ice on the ocean is small since the density of ice is
close to that of fresh water. Hence, the pressure p z=0 is nearly the same whether
there is ice or liquid ocean in the grid cell.
vanishing of grid cells. Additionally, for purposes of simulating the surface mixed
layer, modelers refine the vertical resolution near the surface, and so the depression
of the surface by ice could easily cause the surface grid cell to vanish.
This limitation of z-coordinate models is fundamental. It warrants the consider-
ation of other vertical coordinates. One which has been proposed is p − p atmosphere .
This coordinate will always vanish at the ocean surface, no matter how low the
surface height is. It is presently a research topic to see whether this coordinate is
indeed useful for global ocean climate modeling.
11.7.3 Zero fresh water fluxes and nonzero virtual salt fluxes
Another alternative is to let the ice thickness be unlimited, yet to remove the direct
effects of fresh water forcing from the ocean. That is, we solve the free surface
equation as
η,t = −∇ · U, (11.20)
instead of the budget (11.12). In turn, there is no pressure applied to the ocean from
the overlying ice. It is as if the ice has a zero mass and it exchanges zero mass of
liquid with the ocean. Since the non-tidal dynamical effects yield a surface height
deviation on the order of a few meters, this choice ensures that the ocean model
11.7. THE ALTERNATIVES 143
will not loose surface grid cells, so long at ∆z is something larger than some few
meters (e.g., ∆z = 5m should be sufficient).
By removing fresh water from the volume budget, we reduce the Boussinesq
ocean to an ocean with constant total volume. There is no longer a transfer of vol-
ume across the ocean surface. As the volume transfer between sea ice and ocean is
the problem, one may wish to remove just this transfer, but still allow the ocean to
feel the effects of fresh water from rivers, precipitation, and evaporation. However,
doing so will compromise the conservation of fresh water in the fully coupled sys-
tem. The reason is that ice grows via precipitation (snow). If we allow precipitation
to fall into the ocean as rain, but do not allow fresh water melt from sea ice to enter
the ocean, we have an open loop and so cannot conserve.
Removing fresh water transfers between the ocean and the rest of the climate
system is not the full story for the ocean. The reason is that freshening changes the
ocean salinity, which in turn changes the ocean density. This effect is crucial for
driving ocean currents, and so must be parameterized. As in the rigid lid models,
a means for incorporating this effect into the constant volume ocean model is to
introduce a salt flux across the model surface whenever there is a transfer of fresh
water in the real system.
To motivate the form of the salt flux, let us first consider the case of an ocean
model consisting of a single free surface grid cell affected only by surface water
fluxes. Due to the large hydration energy of salts, there is no salt flux across the
ocean surface, thus leading to the salt budget
That is, total salt is constant in time. Solving for the salinity tendency leads to
s qw
s,t = − , (11.22)
∆z + η
where η,t = qw was used since we are considering the whole ocean to be one grid
cell, and so the ∇ · U term is absent. This budget says that as water is added to the
ocean (qw > 0) it drives salinity to smaller values (s,t < 0).
We now consider the case where there is no transfer of volume across the ocean
surface, and so the total ocean volume is constant. For the present example of
a single grid cell ocean, we therefore have η constant in time. To allow for the
freshening effect, fresh water transported into the ocean is replaced by a virtual salt
flux out of the ocean. In order to provide the same salinity tendency as in the case
of fresh water transfer, we are motivated to define our virtual salt flux as
F exact = −s qw . (11.23)
However, this flux, which is the product of the local time dependent value of salin-
ity and the local fresh water flux, will not allow for conservation of total salt, even
fresh water is globally constant. That is, dx dyF propose is not constant, even if
R
if
R
dx dy qw is constant. Instead, to conserve total salt we must take the virtual salt
flux as
F virtual = −so qw , (11.24)
144 CHAPTER 11. CONSIDERATIONS FOR ICE-OCEAN MODELING
Comparison with the salinity tendency for the case of real fresh water fluxes reveals
that a virtual salt flux yields an accurate salinity tendency only when the constant
salinity so is close to the local salinity s. Errors scale as so /s. With global salinity
running on average near 35psu, it is typical to set so = 35psu. With this value, how-
ever, river mouths introduce problems with this approach, especially with higher
resolution ocean models where the local salinity can become quite small. In this
case, as the river introduces fresh water to the ocean, the virtual salt flux is too
large in magnitude, thus causing the salinity tendency to be too large. Such large
jumps in salinity can introduce numerical problems on top of the physical problems
associated with spuriously large haline forcing in the model.
This result assumes that fresh water transferred across the ocean-atmosphere in-
terface, either as precipitation or evaporation, has the same temperature as the sea
surface temperature carried by the ocean model. With ρ f resh C p ≈ 4 × 106 J/(m3 K ◦ ),
qw on the order of a meter per year, and Twarm − Tcold ≈ 10K ◦ , we have ∆H / A ≈
11.8. HEAT BUDGET IN COUPLED MODELS 145
Contents
12.1 Introduction
Coupling rivers to an ocean model is necessary when building fully coupled climate
models. For z-models, river discharge is typically given fully to the top model grid
cell. Depending on the model resolution, dumping all the river properties to the top
grid cell can cause problems. Notably, without enhanced mixing, a strong halocline
can arise, with associated problems appearing due to noise from vertical advection
across the strong front. This problem is enhanced in models with relatively fine
vertical resolution, such as the 10m now common for the top grid cell in ocean
climate models.
In the real world, there are two reasons that the halocline at river mouths is
somewhat weaker than can occur in ocean climate models. First, river water does
not generally fill only a single layer of some 10m depth. Instead, rivers discharge
into the ocean over a vertical column whose depth can be deeper than 10m. Second,
and more generally, river properties are mixed through a vertical column due to
waves and tides near the coasts.
148 CHAPTER 12. RIVER DISCHARGE INTO THE OCEAN MODEL
Two methods to relieve numerical problems can be considered. First, we can en-
hance vertical mixing of tracers in the region next to river mouths. This approach
is straightforward and is available in MOM4. In detail, the enhanced mixing is
strongest near the surface and tapers to zero at a specified depth. Such is the only
method available to rigid lid ocean models for handling enhanced mixing at river
mouths. Another method is to distribute the river water, along with its tracer con-
tent, over a pre-defined vertical column. Since the top model grid cell in the z-
coordinate MOM4 is the only one capable of changing its volume through changes
in the surface height, distributing river water into deeper cells must be done care-
fully. In the remainder of this chapter, we detail such a method.
Note that we typically do not alter the transfer of momentum from the river
to the ocean. Instead, we assume that river horizontal momentum is the same as
the corresponding ocean cell, thus leading to no change in the ocean momentum
associated with river discharge. This assumption may require modifications for
careful studies of coastal processes, but it should be sufficient for ocean climate
modeling.
Allow the river to be discharging at a volume per area per time given by R, which
has units of a velocity:
The river water flux R is distinguished in MOM4 from fresh water associated with
evaporation and precipitation.
The tracer concentration within the river water is given by Criver :
Criver is distinguished in MOM4 from the tracer concentration associated with evap-
oration and precipitation. What tracer concentration should be taken for the river
water? Typically, we think of rivers at their discharge point as having tracers of uni-
form concentration. More information about tracer profiles requires a river model,
and even so we may wish to summarize the river information prior to passing it
into the ocean. Assuming a single uniform value for the river tracer concentrations,
and absent a river model, it is typical to assume the following river tracer concen-
tration
k=1
θriver = θocean (12.4)
sriver = 0 (12.5)
k=1
Triver = Tocean (12.6)
12.3. STEPS IN THE ALGORITHM 149
where k = 1 is the top cell of the ocean column into which the river water is dis-
charged, θ is the potential temperature, which equals the in situ temperature at the
ocean surface, sriver is the zero salinity of the fresh water river, and Triver is the con-
centration of a passive tracer in the river. By assuming θriver = θocean k=1 , vertically
distributing river water acts to warm the ocean column in regions where the ocean
surface is warmer than depth. In contrast, rivers with zero salinity do not alter the
ocean salt content, yet they do reduce the salinity.
Over a leap-frog tracer time step 2 dtts, a thickness Hriver = R ∗ 2 dtts of river
water is to be distributed throughout the vertical ocean column:
Hriver = R ∗ 2 dtts = river water thickness discharged per tracer leap-frog. (12.7)
Along with this distribution of river water into the ocean column, we distribute the
tracer content of the river into the ocean column:
Criver Hriver = river tracer content discharged per tracer leap-frog. (12.8)
∂t (V ρ C )k=1 = ρo A R Criver
(12.9)
∂ t (V ρ C ) k > 1 = 0
where V = A h is the volume of a grid cell, C is the tracer concentration, ρ is the in
situ density, R is the river discharge rate, and Criver is the concentration of tracer in
the river. As the horizontal area A is constant, it can be dropped from the discus-
sion. Conservation of total tracer in the four-box system is manifest by
kr
∂t ∑ (h ρ C) = ρo R Criver . (12.10)
k=1
Whatever is done to redistribute river runoff with depth, this conservation law
must be preserved.1
To derive the algorithm, we refer to Figure 12.1. Here, we prescribe that a frac-
tion of the river water and its tracer content is inserted into each of the cells within
the column, where the fractions sum to unity
kr
∑ δk = 1. (12.11)
k=1
k=1
δ1 R C water
δ 2R C 2
k=2
δ2 R C water
δ 3 R C3
k=3 δ R C water
3
δ4 R C
4
k=4 δ4 R C water
Figure 12.1: Schematic of river discharge algorithm for the case with kr = 4. We
insert a fraction of the river water into grid cells throughout the column, with a
corresponding amount leaving each cell bubbling upwards in order to conserve
water mass/volume.
where hk is the tracer cell thickness, and is known as dhtk in the MOM4 code.
Because the interior cell volumes remain constant in MOM4, the same amount
of water that entered the cell via the river water must then leave. We assume that
it leaves with the tracer concentration of the cell prior to the insertion of the river
water. That is, by inserting some of the river water into the cell at tracer concen-
tration Criver , we then displace the same amount of water but at concentration Ck .
This displaced water is bubbled upwards towards the surface cell. Conservation
equations for this algorithm take the form
∂t (h ρ C )1 = R [(δ ρ C )2 + δ1 ρo Criver ]
∂t (h ρ C )k = R [(δ ρ C )k+1 − (δ ρ C )k + δk ρo Criver ] (12.13)
∂t (h ρ C )kr = R [−(δ ρ C )kr + δkr ρo Criver ]
where the first equation is for k = 1, the second for 1 < k < kr, and the third
for k = kr. The algorithm has the appearance of upwind advection throughout
the column. Hence, conservation of total tracer for the column is trivially verified.
12.3. STEPS IN THE ALGORITHM 151
For kr = 1, the method reduces to the default discharge of river into the top cell.
152 CHAPTER 12. RIVER DISCHARGE INTO THE OCEAN MODEL
Part V
Quasi-physical Parameterizations
155
Cross-land mixing
Contents
The purpose of this chapter is to present the method used in MOM4 for mixing
tracers and mass/volume across land separated points, such as across an unre-
solved Strait of Gibraltar.
13.1 Introduction
In climate modeling, it is often necessary to allow water masses that are separated
by land to exchange properties. This situation arises in models when the grid mesh
is too coarse to resolve narrow passageways that in reality provide crucial connec-
tions between water masses. For example, coarse grid spacing typically closes off
158 CHAPTER 13. CROSS-LAND MIXING
the Mediterranean from the Atlantic at the Straits of Gibraltar. In this case, it is im-
portant for climate models to include the effects of salty water entering the Atlantic
from the Mediterranean. Likewise, it is important for the Mediterranean to replen-
ish its supply of water from the Atlantic to balance the net evaporation occurring
over the Mediterranean region.
We describe here a method used in MOM4 to establish communication between
bodies of water separated by land. The communication consists of mixing trac-
ers and mass/volume between non-adjacent water columns. Momentum is not
mixed. The scheme conserves total tracer content, total mass or volume (depend-
ing on whether using the non-Boussinesq or Boussinesq versions of MOM4), and
maintains compatibility between the tracer and mass/volume budgets. It’s only re-
striction is that no mixing occur between cells if their time independent thicknesses
differ. This constraint is of little practical consequence.
Notably, mass conservation can be considered a special case of total tracer conser-
vation when the tracer concentration is uniform and constant: T ≡ 1. This re-
sult provides an important compatibility constraint between the discrete tracer and
mass/volume budgets. For constant volume boxes with a Boussinesq fluid, such as
considered in rigid lid models, compatibility is trivial. For boxes which change in
time, such as the top cells in MOM4’s free surface, and/or for non-Boussinesq for-
mulation where total mass is conserved, then compatibility provides an important
constraint on the methods used to discretize the budgets for mass/volume and
tracer. The remainder of this chapter incorporates these ideas into the proposed
cross-land mixing scheme.
An example is the mixing between two constant volume grid cells. If the mixing
takes place instantaneously and between the full contents of both boxes, as in con-
vective adjustment, then the final tracer concentration in both boxes is given by
It is assumed in convective mixing that the volumes of the two boxes remains un-
changed. The picture is of an equal volume of water rapidly mixing from one box
to the other, without any net transport between the boxes.
Instead of instantaneous and complete convective mixing, consider mixing of
the two boxes at a volume rate U. That is, U represents an equal volume per time
of water mixing between the boxes, with no net transport. As shown in Figure 13.1,
U is chosen based on the observed amount of water exchanged through the pas-
sageway. Just as for convective adjustment, the volumes of the two boxes remains
fixed. But the tracer concentrations now have a time tendency. One form for this
tendency relevant for constant volume cells is given by
∂ t (V ( 1 ) T ( 1 ) ) = U ( T ( 2 ) − T ( 1 ) ) (13.5)
∂ t (V (2)
T (2)
) = U (T (1)
−T (2)
). (13.6)
Since the volumes are constant, we can write these budgets in the form
U
∂t T (1) = ( T (2) − T (1) ) (13.7)
V (1)
U
∂t T (2) = ( T (1) − T (2) ) , (13.8)
V (2)
This is the form of cross-land tracer mixing used in the rigid lid full cell MOM1.
In the real world, transport is often comprised of stacked flows where deep wa-
ter flows one way and shallow water oppositely (e.g., see Figure 13.1). Hence, a
more refined form of cross-land mixing may consist of upwind advective fluxes
acting between non-local points in the model, where the advective velocity is spec-
ified based on observations. Such sophistication, however, is not implemented in
MOM4. Indeed, it is arguable that one may not wish to have more details than
provided by the simpler form above, since more details also further constrain the
solution.
ktop ktop
U2
U1
kbot kbot
Figure 13.1: Schematic of cross-land mixing. The model’s grid mesh is assumed
too coarse to explicitly represent the lateral exchange of water masses. For this
schematic, we consider an observed sub-grid scale transport U1 moving in one di-
rection, and U2 in another. To represent the mixing effects on tracers by these trans-
ports, we suggest taking the exchange rate U in MOM4’s cross-land mixing to be
the average of the transports U = (U1 + U2 )/2. Cross-land mixing occurs between
the user-specified depth levels k = ktop and k = kbot . If ktop = 1, then cross-land
mixing of volume in the top cell must be considered, in addition to tracer transport,
in order to maintain compatibility between volume and tracer budgets.
where γ (1) and γ (2) are inverse damping times. This proposed mixing results in
a transfer of mass only when the mass per area within the two boxes differs. The
13.4. MIXING OF MASS/VOLUME 161
total mass of the two-box system is conserved if the following constraint is satisfied
∂t [(ρ h A)(1) + (ρ h A)(2) ] = ( A(1) γ (1) − A(2) γ (2) ) (ρ(2) h(2) − ρ(1) h(1) ) = 0. (13.12)
This relation places a constraint on the inverse damping times γ (1) and γ (2)
13.4.3 A finite time incomplete mixing only when surface heights differ
The following prescription satisfies our desires to mix only when the surface heights
differ
∂t (ρ(1) h(1) ) = γ (1) ρ (h(2) − h(1) ) (13.18)
∂t (ρ(2) h(2) ) = γ (2) ρ (h(1) − h(2) ). (13.19)
When considered over interior time independent model grid cells, then we restrict
its application to grid cells of equal thicknesses. Restricted as such, the right hand
side vanishes for the time independent interior cells, thus leading to no density
mixing. The density factor ρ can be given by anything convenient, such as
ρ(1) + ρ(2)
ρ= , (13.20)
2
or the even simpler prescription (used in MOM4)
2ρ = ρo . (13.21)
162 CHAPTER 13. CROSS-LAND MIXING
• In the rigid lid Boussinesq full cell case, the tracer tendency reduces to equa-
tions (13.7) and (13.8) used in MOM1.
• If the tracer concentration in the two boxes is the same yet the mass differs,
then mixing of mass will leave the tracer concentrations unchanged.
The last constraint is satisfied if the tracer and mass budgets are compatible, as
described in Section 13.2.
Mixing that satisfies these constraints is given by
2Uρ
∂t (ρ (1) (1)
h T (1)
) = ( h(2) T (2) − h(1) T (1) ) (13.22)
A(1) ( h(1) + h(2) )
2Uρ
∂t (ρ(2) h(2) T (2) ) = ( h(1) T (1) − h(2) T (2) ) (13.23)
A(2) ( h(1) + h(2) )
2Uρ
∂t (ρ(1) h(1) ) = ( h(2) − h(1) ) (13.24)
A(1) ( h(1) + h(2) )
2Uρ
∂t (ρ(2) h(2) ) = ( h(1) − h(2) ) (13.25)
A(2) ( h(1) + h(2) )
In these equations, h is the vertical thickness of a tracer cell. For the top cell, thick-
ness takes the form
h = ∆z + η (13.26)
where ∆z is the time independent thickness of the fixed volume rigid lid case, and
η is the surface height. The mass per area equations (13.24) and (13.25) result from
the tracer equations (13.22) and (13.23) upon setting the tracer concentrations to a
constant, as required for compatible budgets. When the cell thicknesses are time
independent, such as the ocean model interior with k > 1, then we restrict our
applications to cases where h(1) = h(2) , for reasons articulated in Section 13.4.2.
where
(lx) (lx)
A(lx) = dxti, j dyti, j (13.28)
are the generally different horizontal cross-sectional areas of the tracer cells in the
two columns, and H (1) = H (2) is the vertical thickness of the two columns. The top
and bottom k-levels for the columns are set by k = ktop and k = kbot. As mentioned
earlier, the formulation here allows for mixing only between boxes that live on the
same k-level, so k = ktop and k = kbot are the same for both columns lx = 1, 2.
Use of these volumes in equations (13.7) and (13.8) leads to the tracer time ten-
dencies for a particular k-level
(1) (2) (1)
∂t Tk = B(1) ( Tk − Tk ) (13.29)
(2) (1) (2)
∂t Tk = B(2) ( Tk − Tk ), (13.30)
where
U
B(lx) = (13.31)
V (lx)
represents the rate (B(lx) has units of inverse time) at which the two columns par-
ticipate in the mixing. Conservation of total tracer is maintained between two hor-
izontally adjacent boxes within the two columns. We see such conservation via
multiplying the above tendencies by the respective time independent volumes of
the two cells, and adding
(1) (1) (2) (2) (2) (1) (1) (2)
∂t (Vk Tk + Vk Tk ) = ( Tk − Tk ) ( A(1) B(1) hk − A(2) B(2) hk ) = 0, (13.32)
(1) (1) (2) (2)
where A(1) B(1) hk = U (hk / H (1) ) = U (hk / H (2) ) = A(2) B(2) hk was used.
where H (1) and H (2) are the generally different depths of the two columns. By
inspection, for each k-level this formulation conserves total tracer mass and total
fluid mass (recall Section 13.4). Also, when the mass per area (ρ h) is constant in
time, as in a rigid lid Boussinesq fluid, and when H (1) = H (2) , as in a full cell
model, then these budgets reduce to equations (13.29) and (13.30) used in MOM1.
Furthermore, setting the tracers to uniform constants leads to the transfer of mass
per area between two cells living at the same k-level
(1) (1) 2 U ρk (2) (1)
∂t (ρk hk ) = ( hk − hk ) (13.35)
A(1) ( H (1) + H (2) )
(2) (2) 2 U ρk (1) (2)
∂t (ρk hk ) = ( hk − hk ). (13.36)
A(2) ( H (1) + H (2) )
Again, for time independent cells with k > 1, we restrict our applications to those
cells with h(1) = h(2) , which means there is no added density mixing. This restric-
tion is minimal as it only precludes our using cross-land mixing within a partial cell
bottom.
Both of these budgets can be written in a form familiar from other damping
processes
Just as for any other form of mixing, if the damping coefficients are too large, then
it is possible for there to be numerical instabilities. MOM4 provides a check so that
no more than one-half of a particular grid cell is mixed per model time step.
In this equation, ρk hk is the mass per unit area within a tracer grid cell with vertical
label k, −ρo ∇ · U is the convergence of the vertically integrated horizontal velocity
with U = ∑kNk =1 hk uk , and ρw qw is the mass per horizontal area per time of fresh
water entering the ocean surface. In words, this budget says that the mass per
horizontal area contained in a column of fluid changes in time according to the
convergence of the vertically integrated horizontal mass per area, plus any mass
per area entering the ocean through its surface. Since only the top model grid cell
has a time-dependent thickness
h1 ( x, y, t) = η( x, y, t) + ∆z, (13.46)
we can use the mass balance over a column to derive a prognostic equation for
surface height
Nk
ρ1 η,t = −ρo ∇ · U + ρw qw − ∑ hk ∂t ρk . (13.47)
k=1
The density time derivative term accounts for the so-called steric effects which arise
from time changes in the vertically integrated density. This term vanishes for a
Boussinesq fluid.
The semi-discrete budget for tracer within a discrete grid cell is given by
∂t (ρ h T ) = −ρo ∇ · (h F) − ρo h ∂ z F z , (13.48)
where the equations are written for the point (1) and δk1 is unity for k = 1 and
vanishes elsewhere. Again, the added mass source is present only for the surface
cell since we restrict application to those cases where the time independent interior
cells have h(2) = h(1) .
Given these results, we identify the density weighted source for surface height
All terms on the left hand side are evaluated in time at the present time step τ.
Those terms on the right hand side are evaluated at the lagged time τ − 1, as is
required for linear numerical stability. The density and thickness weighted tracer
concentration source takes the form
where again terms on the left hand side are evaluated at time τ whereas those on
the right are at time τ − 1.
In practice, the tracer equation is time stepped for tracer concentration, with the
time tendencies h,t and ρ,t forcing tracer concentration tendencies. In this case, we
consider
where h,t vanishes for interior cells with k > 1. Care should be taken to include
the contributions to h,t arising from all processes, including cross-land mixing. For
example, if cell (1) is in the surface k = 1, then the tracer concentration budget for
this cell takes the form
where
Nk
ρ1 h,tno−xland = −ρo ∇ · U + ρw qw − ∑ hk ∂t ρk (13.56)
k=1
13.8. SUPPRESSION OF B-GRID NULL MODE 167
is the time tendency for the surface height, sans the cross-land mixing effects (see
the mass conservation equation (13.45)). Time indices on the cross-land terms are
given by
Contents
The purpose of this chapter is to detail the method for diffusing tracers along
the bottom topography within the bottom-most model grid cells.
14.2 Diffusivities
When applying sigma-diffusion, the diffusive flux between two adjacent cells living
at the ocean bottom is given by
Fσ = − A ∇σ T, (14.1)
with ∇σ the horizontal gradient operator taken between cells in the bottom sigma-
layer. We follow the approach of Döscher and Beckmann (1999) in which sigma-
diffusion dominates when favorable according to densities of the participating cells.
That is, the following diffusivity is used
Amax if ∇σ ρ · ∇ H < 0
A= (14.2)
Amin if ∇σ ρ · ∇ H ≥ 0,
In this expression, |u| is the magnitude of the zonal velocity component and ∆x is
the zonal grid spacing. An analogous meridional flux is computed as well.
14.3 Implementation
The bottom sigma-layer in MOM3 was appended to the very bottom of the model
and effectively lived beneath the deepest rock. This approach has its advantages.
However, it makes for awkward analyses; it precludes direct comparison between
models run with and without sigma-physics since the grid used by the respective
models is different; and it makes it difficult to consider convergence when refin-
ing the grid mesh. For these reasons, the bottom sigma-layer in MOM4 is included
along with the rest of the model domain. This is the approach used by Beckmann
and Döscher (1997) (e.g., see their Figures 1 and 2). The disadvantage is that the
sigma-layer thickness bbl thickness in MOM4 has a generally non-constant thick-
ness determined by the thickness of the model grid cell next to topography.
To address the limitation of having bbl thickness constrained by the bottom par-
tial cell thickness dht, we suggest two approaches. For the simple case where dht is
thick, then we set
where sigma thickness max is a namelist parameter. Values on the order of 50m are
suggested.
14.3. IMPLEMENTATION 171
vert
σ horz
σ
σ
Figure 14.1: Schematic of the extra along-topography pathway for diffusive trans-
port afforded by the sigma-diffusion scheme applied to the bottom-most model
cells. Darkened regions denote land cells. Tracers in the z-level cells communicate
with their horizontal and vertical neighbors via the usual advection, diffusion, and
convective processes. Tracers in the bottom “sigma-layer” can also communicate
with their sigma-neighbors via sigma-diffusion.
For the more complicated case where dht is thin, then we have no work-around
since doing so involves allowing the sigma-diffusion process to take place between
arbitrary numbers of grid cells laterally adjacent within the sigma layer. The only
option at present is to limit the partial cell thicknesses dht to be no smaller than
ones desired minimum bbl thickness. This approach reduces the fidelity of the bot-
tom topography representation, but perhaps by gaining a better transport by sigma
diffusion. Such issues remain an ongoing research topic at GFDL.
Consistent with Beckmann and Döscher, the sigma-layer momentum equations
remain the same as interior z-level cells. We now just allow tracers in the bottom-
most cell to be affected by diffusion with their “sigma-neighbors” in addition to
their horizontal and vertical neighbors. If desired, communication with horizontal
and vertical neighbors can be removed for the sigma cells. Figure 14.1 provides a
schematic of the extra sigma-pathway available with sigma tracer diffusion.
172 CHAPTER 14. SIGMA TRACER DIFFUSION
CHAPTER
FIFTEEN
Contents
15.1 The ubiquitous cliffs in coarse z-models . . . . . . . . . . . . . . 173
15.2 The Campin and Goosse (1999) algorithm . . . . . . . . . . . . . 174
15.2.1 Finding the depth of neutral buoyancy . . . . . . . . . . . 174
15.2.2 Prescribing the downslope flow . . . . . . . . . . . . . . . 175
15.2.3 Mass conservation and tracer transport . . . . . . . . . . . 176
15.3 Implementation in MOM4 . . . . . . . . . . . . . . . . . . . . . . 177
15.3.1 Start of the integration . . . . . . . . . . . . . . . . . . . . . 177
15.3.2 During a time step . . . . . . . . . . . . . . . . . . . . . . . 178
15.4 Comments on the two overflow schemes in MOM4 . . . . . . . . 179
This chapter documents the MOM4 implementation of the Campin and Goosse
(1999) scheme that discharges dense overflow waters into the deep. The goal of the
scheme is similar to that of the sigma-diffusion scheme of Beckmann and Döscher
(1997), as described in Chapter 14. We discuss differences between the two methods
in Section 15.4.
than the typical resolution of the 1-2 degree ocean climate models commonly used
today. Note that refined vertical resolution, desired for resolving/parameterizating
vertical physical processes, requires one to further refine the horizontal resolution
required to resolve the slope. Notably, there is little difference between the rep-
resentation of steeply sloping features via either full or partial cells in z-models.
Hence, such cliff features remain ubiquitous in the typical ocean climate z-model.
It due to our inability to fully resolve such steeply sloping topographic features
that we require enhanced bottom mixing/transport whose aim is to allow dense
water to move into the deep, rather than sitting on top of the shelf.
ρso (kup, kup) = ρ(si, j,kup , θi, j,kup , pi, j,kup ) (15.2)
so
ρ (kup, kdw − 1) = ρ(si, j,kup , θi, j,kup , pi+1, j,kdw−1 ) (15.3)
ρso (kup, kdw) = ρ(si, j,kup , θi, j,kup , pi+1, j,kdw ) (15.4)
so
ρ (kup, kdw + 1) = ρ(si, j,kup , θi, j,kup , pi+1, j,kdw+1 ). (15.5)
15.2. THE CAMPIN AND GOOSSE (1999) ALGORITHM 175
do so do
3 ρ(kdw−1) ρ(kup,kdw−1) > ρ(kdw−1)
i,j
do
ρ(kdw) do
so
k 4 ρ(kup,kdw) > ρ(kdw )
so do
5
do
ρ(kdw+1) ρ(kup,kdw+1) < ρ(kdw+1)
Figure 15.1: Schematic of the Campin and Goosse (1999) overflow method in the
horizontal-vertical plane. The darkly filled region represents bottom topography
using MOM4’s full cells. The lightly filled region represents topography filled by a
partial cell. Generally, the thickness of a cell sitting on top of a topographic feature,
as the k = 2 cell in the “so” column, is thinner than the corresponding cell in the
deep-ocean column (the k = 2 cell in the “do” column). Shown are tracer cells, with
arrows representing the sense of the scheme’s upstream advective transport. This
figure is based on Figure 1 of Campin and Goosse (1999).
That is, we compute the density at the salinity and potential temperature of the
shallow ocean parcel, (si, j,kup , θi, j,kup ), but at the in situ pressure for the respective
grid cell in the deep column. The density is then compared to the density of the
parcel at the in situ salinity, temperature, and pressure of the cells in the deep ocean
column.
where
V (t) = dxt ∗ dyt ∗ dht (15.9)
is the volume of the dense parcel’s tracer cell. Equation (15.8) is also used to de-
termine a meridionally directed downslope transport, with the meridional topo-
graphic slope H,y replacing H,x , and ∆ρ the density difference between meridion-
ally adjacent parcels.
Solving equation (15.8) for the speed uslope yields
slope gδ
u = | H,x | ∆ρ. (15.10)
ρo µ
If the depth H refers to the depth of a tracer cell, then the absolute slope | H,x | is
naturally defined at the zonal face of the tracer cell. Hence, the speed, uslope , is
likewise positioned at the zonal face. This is the desired position for an advective
tracer transport velocity.
Campin and Goosse (1999) suggest the values µ = 10−4 sec−1 and δ = 1/3.
These parameters are set as namelists in MOM4. Using these numbers, with an
absolute topographic slope of | H,x | ≈ 10−3 and density difference ∆ρ ≈ 1 kg m−3 ,
leads to the speed
uslope ≈ .03 m sec−1 . (15.11)
Associated with this downslope speed is a volume transport of fluid leaving the
cell
U slope = uslope dhtmin dyt. (15.12)
In this equation, dhtmin is the minimum thickness of the shelf cell and the adjacent
cell. This minimum operation is necessary when considering MOM4’s bottom par-
tial cells, whereby the bottom-most cell in a column can have arbitrary thickness
(Figure 15.1). With uslope ≈ .03m s−1 corresponding to the speed of fluid leaving a
grid cell that is one-degree in width and 50 m in depth, we have a volume transport
U slope ≈ 0.2Sv. Larger values are easily realized for steeper slopes, larger density
differences, and larger grid cells.
The convergence-free seawater mass flux carries with it tracer mass. If there are
differences in the tracer content of the cells, then the tracer flux will have a nonzero
convergence, and so it moves tracer throughout the system. We use first-order up-
stream advective transport as a discretization of this process. First-order upstream
advection is the simplest form of advection. Its large level of numerical diffusion is
consistent with our belief that the bottom layer flows in the real ocean near steep
topography are quite turbulent. Hence, although inappropriate for interior flows,
we are satisfied with the use of upstream advection for the overflow scheme.
(i,j,2)
i
(i,j,4)
Figure 15.2: Plan view (x-y plane) of a tracer grid cell at (i, j) and its horizontally
adjacent tracer cells. We label the adjacent cells (i + 1, j), (i, j + 1), (i − 1, j), (i, j − 1)
as m = 1, 2, 3, 4. Notice that we do not consider downslope flow along a diagonal
direction.
178 CHAPTER 15. DISCHARGING OVERFLOW WATERS INTO THE DEEP
where
A(t) = dxt ∗ dyt (15.22)
is the time independent horizontal area of a tracer cell.
ρ=1035
1
ρ=1033
2 ρ=1034
i,j
ρ=1035
k 3
4 ρ=1036
5
ρ=1037
Figure 15.3: Schematic of a situation where a dense parcel sits on a shelf next to a
column whose upper portion is light, but whose deeper portion is denser than the
shelf. For this case, the Campin and Goosse (1999) scheme prescribes a transport
between the shelf water at level 1 and the deeper water at level 3, with water bub-
bling upward to conserve mass as shown in Figure 15.1. In contrast, the Beckmann
and Döscher (1997) scheme will not prescribe any enhanced transport, since here
the bottom of the deep column is denser than the shelf.
Part VI
Some diagnostics
183
S OME DIAGNOSTICS
MOM4 makes extensive use of the FMS diagnostic manager, whereby any field
that is carried by the model can be trivially output to a diagnostics file via the reg-
istration of that field with the diagnostics manager, and the associated entry of the
field’s name within the diagnostics table. This part of the MOM4 Guide does not
describe this aspect of diagnostics. Instead, we focus here on the formulation and
computation of certain diagnostics that have proven of use for research with z-
coordinate models.
184
CHAPTER
SIXTEEN
Streamfunctions
Contents
ρ,t + ρo ∇ · v = 0. (16.1)
186 CHAPTER 16. STREAMFUNCTIONS
In orthogonal horizontal coordinates, velocity field divergence takes the form (see
Griffies (2004))
∇ · v = (dy dz)−1 (dy dz u),x + (dx dz)−1 (dx dz v),y + (dx dy)−1 (dx dy w),z
(16.2)
which then leads to the continuity equation
(dV /ρo ) ρ,t + dx (dy dz u),x + dy (dx dz v),y + dz (dx dy w),z = 0, (16.3)
where
dV = dx dy dz (16.4)
corresponds to the volume of a model grid cell. Note that the volume of grid cells in
the top levels of MOM4 are time dependent due to the use of a free surface. Hence,
it is not generally valid to bring dV inside the time derivative.
To set the notation, the distance between two infinitesimally close points on the
sphere is written
1
Zξb
V = (dz)−1 dx (v dz) (16.6)
ξ a1
1
Zξb
W = (dy)−1 dx (w dy). (16.7)
ξ a1
ξ a1
ξb1
1
Z
= ∂2 (dx dz v)
h2
ξ a1
1
Zξb
1
= dy (dx dz v),y (16.8)
dy
ξ a1
These steps made use of the properties of orthogonal coordinates whereby ∂2 com-
mutes with the zonal integration, since it is orthogonal to the integration pathway
taken along a line of constant ξ 1 . However, ∂2 does not commute with the distances
dx and dz, each of which can be functions of the generalized meridional coordinate
ξ 2 . Use of continuity in the form (16.3) yields
1
Zξb
− dy (dz V ),y = [dz (dx dy w),z + (dV /ρo ) ρ,t ]
ξ a1
1
Zξb
= dz (dy W ),z + (dV /ρo ) ρ,t . (16.9)
ξ a1
This relation indicates that meridional and vertical transports are balanced by fluc-
tuations in mass. In a steady state, or for the Boussinesq models, the mass fluctua-
tions vanish, thus leading to volume conservation manifesting as the two-dimensional
non-divergence condition
∂( h3 V ) ∂( h2 W )
2 3
dy (dz V ),y + dz (dy W ),z = dξ dξ + = 0. (16.12)
∂ξ 2 ∂ξ 3
188 CHAPTER 16. STREAMFUNCTIONS
The second equality follows by the orthogonal nature of the coordinates, whereby
∂(dξ 2 ) ∂(dξ 3 )
= = 0. (16.13)
∂ξ 3 ∂ξ 2
for the steady state , or for the Boussinesq models,We are thus motivated to intro-
duce an overturning streamfunction that satisfies
V = −Ψ,z (16.14)
W = Ψ,y . (16.15)
Zz
Ψ( y, z) = Ψ( y, zo ) − dz0 V ( y, z0 ), (16.16)
zo
Zy
Ψ( y, z) = Ψ( yo , z) + dy0 W ( y0 , z), (16.17)
yo
Ry
yo dy0 W ( y0 , zo ) vanishes. For the free surface, however, the possibility of surface
water fluxes allows for an open water budget above the ocean surface. Since there
is no attempt here to account for water cycling through the rock beneath the ocean,
one can assume all water transport in rock vanishes. Hence, by taking zo to be some
value completely beneath the ocean bottom, the vertical transport term can again
be dropped with the free surface. As a consequence, a general expression for the
overturning streamfunction, valid for both the free surface and rigid lid, is given
just by the meridional transport term
Zz Zz Zxb
0 0
Ψ( y, z) = − dz V ( y, z ) = − dx dz0 v. (16.20)
zo zo x a
In practice, it is not necessary to evaluate the integral anyplace beneath the ocean
bottom, since the water velocity vanishes there. On the bottom, the vertical coor-
dinate takes on the non-constant value z = − H ( x, y). Therefore, when integrating
just to the ocean bottom, it is necessary to perform the vertical integral first, and
then the zonal integral
Zxb Zz
Ψ( y, z) = − dx dz0 v. (16.21)
xa −H
Beneath the ocean surface, the no-normal flow condition implies that the over-
turning streamfunction is a constant along the side and bottom land-sea bound-
aries. For the rigid lid, the absence of fresh water input to the ocean surface also
implies that its overturning streamfunction is a constant at the ocean surface. The
choice Ψ( yo , zo ) = 0 means that the rid lid overturning streamfunction is zero along
all the boundaries. For the free surface, however, the overturning streamfunction
need not be a constant on the ocean surface, due to the presence of surface water
fluxes, whereas it remains zero on the sides/bottom just as for the rigid lid.
s( z)
Zxb Z
Ψ( y, s) = − dx dz0 v, (16.22)
x a −s( H )
with s = s( x, y, z, t) the generalized vertical coordinate (see Griffies (2004) for de-
tails). Surfaces that are physically of interest include various potential density sur-
faces, which are especially relevant when the flow is adiabatic. See Section 40.9 of
Pacanowski and Griffies (1999) for more discussion.
190 CHAPTER 16. STREAMFUNCTIONS
where
v gm = −∂ z (κ S y ) (16.24)
is the eddy-induced velocity with S y = −∂ y ρ/∂ z ρ the neutral slope in the y-direction
and κ a diffusivity. Performing the vertical integral on the GM90 piece leads to
Ψ(tot) ( y, z) = Ψ( y, z) + Ψ gm ( y, z) (16.25)
where
Zxb
gm
Ψ ( y, z) = dx (κ S y ) (16.26)
xa
Ψ gm ∼ L S κ. (16.27)
where dl is the line element along any path connecting the points a and b, and n̂ is a
unit vector pointing perpendicular to the path in a rightward direction when facing
16.2. VERTICALLY INTEGRATED TRANSPORT 191
the direction of integration. As written, Tab has units of volume per time, and so it
represents a volume transport. Therefore, the difference between the streamfunc-
tion at two points represents the vertically integrated volume transport between
the two points. It is for this reason that the streamfunction is sometimes called
the volume transport streamfunction. Note that Bryan (1969) defined the barotropic
streamfunction with an extra factor of the Boussinesq density ρo , such that his mass
transport streamfunction has the dimensions of mass per time rather than volume
per time. The difference is trivial for a Boussinesq fluid.
is given. Although accurate and complete, this integral does not readily provide a
horizontal map of transport, and so it looses much of the appeal associated with
the transport streamfunction used with a rigid lid.
However, for many practical situations, maps of the function
Z y
ψ( x, y) = − dy0 U ( x, y0 ) (16.32)
yo
are quite useful, where the lower limit yo is taken at the southern boundary of the
domain, generally given by a solid wall for ocean climate models. By its definition,
the meridional derivative of ψ yields the zonal transport
ψ,y = U. (16.33)
The zonal derivative, however, does not yield the meridional transport due to the
divergent nature of the vertically integrated flow. To see what it does yield, let us
restrict attention to the Boussinesq case with spherical coordinates, where
Z φ
1 R
ψ,λ = V − dφ0 cos φ0 ∇ · U. (16.34)
R cos φ cos φ φo
To reach this result we used where cos φo V (λ, φo ) = 0. With volume conservation
in the Boussinesq model leading to
∇ · U = −ηt + qw , (16.35)
192 CHAPTER 16. STREAMFUNCTIONS
we see that ∇ · U vanishes only when there is zero time tendency of the free surface
height and zero fresh water flux through the surface. Hence, the zonal derivative
leads to the meridional transport plus a term that vanishes in the case of a steady
state (ηt = 0) and a zero fresh water flux. It is useful to see how large the extra term
might be. For this purpose, zonally integrate the derivative ψ,λ to find
Z λ
ψ(λ, φ) − ψ(λo , φ) = R cos φ dλ 0 V (λ 0 , φ)
λo
Z λ Z φ
− R2 dλ 0 dφ0 cos φ0 (−ηt + qw ). (16.36)
λo φo
For example, with −ηt + qw = 1m/ year applied over a 1000km × 1000km area,
the extra term contributes much less than a Sv to the streamfunction. Cases where
the differences are larger certainly can be constructed. But for many diagnostic
purposes, the differences are negligible.
By construction, ψ reduces to the transport streamfunction in the case of a rigid
lid where ∇ · U = 0. However, this is not a unique choice and alternatives do exist.
For example, Z x
ψ∗ ( x, y) = ψ( xo , y) + dx0 V ( x0 , y), (16.37)
xo
gives
V = ψ∗,x . (16.38)
ψ∗ has the advantage that zonal derivatives give the exact meridional transport, yet
the meridional derivative deviates from the zonal transport. In the end, it might be
useful to plot ψ and ψ∗ and compare.
As each streamfunction is defined only up to an arbitrary constant, it is useful
to specify this constant in a manner to correspond to that resulting from the rigid
lid approximation. To do so, it is recommended that one normalizes each stream-
function by the value at λ = 300◦ and φ = −20◦ , which corresponds to a point over
South America. This convention corresponds to taking the Americas as the zeroth
island in the rigid lid method.
CHAPTER
SEVENTEEN
Contents
17.1 Introduction
Diagnosing the transport of tracers across a particular section or through a straight
provides a useful means to compare simulations with observations. It also allows
one to summarize a large amount of information thus integrating together many
processes. As described in various papers, such as Bryan (1987), one common inte-
grated diagnostic is the meridional tracer transport. Furthermore, when coupling
to an atmospheric model, Gordon et al. (2000) found it crucial to roughly match
the ocean meridional heat transport to that expected by the atmospheric model.
Motivated by the importance of meridional tracer transport to climate modeling,
we focus mostly on meridional transport, though all results can be generalized to
arbitrary directions. For the tripolar grid supported by MOM4 (Section 4.2), the fol-
lowing diagnostics represent true northward transport for regions south of 65◦ N.
194 CHAPTER 17. DIAGNOSING TRACER TRANSPORT
In this equation,
Zη
U= dz u = ( H + η) u (17.3)
−H
is the depth integrated velocity, qw is the volume per area per time of fresh water
entering the ocean surface (qw > 0 means water enters the ocean domain), z = − H
is the ocean bottom, and is z = η the free surface deviation from z = 0. Derivation
of equation (17.2) required the use of the surface and bottom kinematic boundary
conditions (see Section 3.4). In a semi-discrete form appropriate for the discrete
numerical model, the column mass budget is given by
!
∂t ∑hρ + ρo ∇ · U = ρo qw (17.4)
k
At steady state, the zonally integrated divergence of the vertically integrated merid-
ional velocity balances the zonally integrated fresh water flux
Zxe Zxe
dx ∂ y V = dx qw steady state. (17.6)
xw xw
At steady state, the zonal and vertical integrated velocity at a particular latitude
balances the zonal and vertical integrated fresh water flux
∂t (ρ h T ) = −ρo ∇ · (h F) − ρo h ∂ z F z , (17.10)
where we drop tracer source terms for simplicity. In this equation, h is the thickness
of the model grid cell, and F = (Fh , F z ) is the combined advective and sub-grid-
scale tracer fluxes
F = v T + Fsgs . (17.11)
Integration over the depth of the ocean and along an i-line (the zonal direction
when using spherical coordinates), and assuming either solid walls or periodic
boundary conditions, renders
!
∂t ∑ dx dy h ρ T = −ρo ∑ δ j (dx h F y ) − ρo ∑ dx dy Fkz=0 , (17.12)
i,k i,k i
where
δ j Φ = Φ j+1 − Φ j (17.13)
is a difference operator acting on the fluxes passing across the meridional faces of
the tracer cells, and we employed the discrete divergence operator introduced in
Section 6.2. The vertical flux at k = 0 is given by (see Griffies (2004))
In the absence of surface tracer fluxes, the steady state flux moving northward must
balance that moving southward. The result reduces to equation (17.8) in the case
196 CHAPTER 17. DIAGNOSING TRACER TRANSPORT
of tracer concentration is uniform. Note that only for coordinates where dy is inde-
pendent of i (such as spherical coordinates) can dy be transferred to the right hand
side.
Now integrate from a southern row j = js where the meridional flux vanishes,
such as the Antarctic continent. At the northern row j = jn, the meridional and
zonal integrated surface flux is balanced at steady state by the net meridional flux
entering southward through the northern j-row boundary
j= jn
∑(dx h F y ) j= jn = − ∑ ∑ dx dy Fkz=0 steady state. (17.16)
i,k i j= js
This result reduces to equation (17.9) in the case of uniform tracer concentration
and zero surface tracer fluxes.
∑ dx h F y = ρo ∑ dx (v h T + h Fsgs ).
y
F north = ρo (17.17)
i,k i,k
where
Zη nk
1 1
Φ=
H+η
dz Φ −→
H+η ∑ hΦ (17.19)
−H k=1
o Fadvect = ∑ dx
ρ− ∑ (v h) T
1 north
(17.20)
i k
= ∑ dx V T + ∑ dx ∑ vch T.
b (17.22)
i i k
17.4. NORTHWARD TRACER TRANSPORT 197
We now decompose fields according to their zonal mean and deviations therefrom
0
Φ = (Φ − Φ x ) + Φ x = Φ x + Φ x (17.23)
where
Zx2
1 1
x
Φ =
L
dx Φ −→
L ∑ dx Φ (17.24)
x1 i
o Fadvect = ∑ dx V T + ∑ dx
ρ− ∑ vch Tb
1 north
i i k
0
x0 x
bx + vch x T
= Lk=0 V T + ∑ dx V x T + ∑ ∑ dx (vch
x x 0
b x0 )
T
i k i
x x x0 x0
= Lk=0 V x T + ∑ dx V x T + ∑ Lk vch T
b +∑ ∑ dx vch
x 0
b x0
T (17.25)
i k k i
where Lk is the distance as a function of vertical depth level over which the zonal
integration takes place within sea water (i.e., land is not included).
The term
north x x
Fmean −mean = ρo Lk=0 V T (17.26)
represents the zonal and depth mean meridional velocity advecting the zonal and
depth mean tracer. In a steady state with no surface forcing, equation (17.9) showed
that mass conservation leads to V x = 0, and so Fmean
north
−mean = 0. In general, however,
this transport is not zero.
The term
x0
∑ dx V x
0
Fbtnorth
− gyre = ρo T (17.27)
i
is called the barotropic gyre transport. The gyre adjective refers to the presence of
zonal anomalies in both the vertically integrated velocity and vertically integrated
tracer. Such anomalies typically arise from wind forcing in ocean basins. This term
vanishes in a domain with zero zonal anomalies and/or with zero vertically inte-
grated flow.
198 CHAPTER 17. DIAGNOSING TRACER TRANSPORT
is called the baroclinic gyre transport. Again, often one extracts from this term an
Ekman component described in Section 17.4.3.
Contents
The purpose of this chapter is to detail a method to quantify water mass mix-
ing in MOM without detailed knowledge of the numerical transport scheme. The
method is restricted to experiments with a Boussinesq fluid, flat bottom ocean, lin-
ear equation of state, and without buoyancy forcing. Extensions are possible, yet
not implemented. Momentum forcing via winds is allowed. Much of the funda-
mentals in this chapter are guided by the work of Winters et al. (1995) and Winters
200 CHAPTER 18. EFFECTIVE DIANEUTRAL DIFFUSIVITY
and D’Asaro (1995). Griffies et al. (2000b) applied these methods to various ideal-
ized model configurations.
We assume the linear equation of state for an incompressible fluid is written in
the form
ρ = ρ0 (1 − α θ ), (18.1)
where θ is potential temperature, ρo is a constant density associated with the Boussi-
nesq approximation, and α is a constant thermal expansion coefficient. The system
is open to momentum fluxes yet closed to buoyancy fluxes.
where
P = gz (18.3)
is the potential energy per mass of a fluid parcel, g is the acceleration of gravity, z
is the vertical position of a fluid parcel, and ρ dV = ρ dx dy dz is the parcel mass.
Available potential energy (APE) is the difference between the potential energy
of the fluid in its natural state, and the potential energy of a corresponding stably
stratified reference state. The reference state is reached by adiabatically rearranging
the fluid to a state of minimum potential energy, which is a state that contains zero
horizontal gradients. This rearrangement, or sorting, provides a non-local mapping
between the unsorted fluid density and the sorted density
The sorting map determines a vertical position field z∗ (x, t) which is the vertical
height in the sorted state occupied by a parcel at (x, t) in the unsorted state. Due
to the monotonic arrangement of density in the sorted state, z∗ (x, t) is a monotonic
function of density ρ(x, t).
It is convenient to set the origin of the vertical coordinate at the ocean bottom so
to keep potential energy of the unsorted state non-negative. This convention also
allows for z∗ (x, t) to be defined as a monotonically decreasing function of density.
That is,
ρ(x1 , t) < ρ(x2 , t) ⇒ z∗ (x1 , t) > z∗ (x2 , t). (18.5)
Conservation of volume in a flat bottom ocean implies that the sorted fluid state
has the same vertical extent as the unsorted fluid, which renders
0 ≤ z, z∗ ≤ H, (18.6)
In the following, it proves convenient to denote the density profile in the sorted
reference state using the symbols
Given this notation, the non-local sorting map between the unsorted and sorted
fluid states provides the equivalence
In turn, potential energy for the sorted fluid state can be written in two equivalent
manners
Z
Ere f = g dV z ρre f ( z, t) (18.9)
Z
= g dV z∗ (x, t) ρ(x, t). (18.10)
Equation (18.9) represents an integral over the sorted fluid state, in which the den-
sity of this state is a function only of the depth. The horizontal area integral is thus
trivial to perform. Equation (18.10) represents an integral over the unsorted fluid
state, where the density ρ(x, t) of an unsorted parcel is weighted by the vertical
position z∗ (x, t) that the parcel occupies in the sorted state. It follows that the APE
can be written in two equivalent ways
Z
E APE = g dV z [ρ(x, t) − ρre f ( z, t)] (18.11)
Z
= g dV ρ(x, t) [ z − z∗ (x, t)]. (18.12)
This result, derived for a closed fluid system, suggests the introduction of a global
effective diffusivity
∂t dV ρre2
R !
f
κ global (t) = − . (18.15)
2 dV (∂ z ρre f )2
R
This diffusivity provides one number that can be used to represent the total amount
of dianeutral diffusion acting over the full model domain. It vanishes when the
simulation is adiabatic, as does the effective diffusivity κe f f ( z∗ , t). However it is
generally different from the vertical average of κe f f ( z∗ , t).
is naturally defined at the top face of the density cell whose center is at z∗ . As such,
the diffusion operator at the lattice point z∗ , which is constructed as the conver-
gence of the diffusive flux across a density grid cell, takes the discrete form
∗ ∗
F z ( z∗ , t − ∆t) − F z ( z∗ − ∆z∗ , t − ∆t)
z∗
− (∂ z∗ F )( z , t) ≈ −
∗
. (18.17)
∆z∗
The time lag is necessary to provide for a stable discretization of the diffusion equa-
tion. The discretization of the flux is given by
∗
F z ( z∗ , t) = −κe f f ( z∗ , t) ∂ z∗ ρre f ( z∗ , t)
ρre f ( z∗ + ∆∗ z, t) − ρre f ( z∗ , t)
∗
≈ −κe f f ( z , t) . (18.18)
∆z∗
18.2. EFFECTIVE DIANEUTRAL MIXING 203
Since the flux is located at the top face of the density grid cell whose center is at
the position z∗ , the effective diffusivity is located at this face as well. Each of these
difference operators is consistent with those used in MOM when discretizing the
diffusion equation for the unsorted fluid.
As with the unsorted tendency, the time derivative in the effective diffusion
equation can be approximated using a leap-frog differencing:
ρre f ( z∗ , t + ∆t) − ρre f ( z∗ , t − ∆t)
∂t ρre f ( z∗ , t) ≈ . (18.19)
2∆t
Piecing these results together yields the expression for the vertical flux at the
top of the density cell z∗ + ∆z∗
∗ ∗ ∆z∗
F z ( z∗ , t − ∆t) = F z ( z∗ − ∆z∗ , t − ∆t) − [ρ ( z∗ , t + ∆t) − ρre f ( z∗ , t − ∆t)].
2∆t re f
(18.20)
This flux can be determined starting from the ocean bottom, where is vanishes, and
working upwards. Without surface buoyancy fluxes, it also vanishes at the top of
∗
the water column, resulting in conservation of dz ρre f ( z∗ , t). After diagnosing
R
the flux from the tendency, the effective diffusivity can be diagnosed from
∆z∗
∗
κe f f ( z∗ , t) = − F z ( z∗ , t) . (18.21)
ρre f ( z∗ + ∆z∗ , t) − ρre f (∗ z, t)
The issues of what to do when the density gradient becomes small, as in weakly
stratified regions, is discussed in Sections 18.2.3 and 18.2.5.
g dρre f
N∗2 = −
ρo dz∗
g dσre f
= − , (18.22)
1000 ρo dz∗
where σre f = 1000 (ρre f − 1) is the sigma value for the sorted density ρre f ( g/cm3 ).
Working with σre f is desirable for accuracy reasons. The observed range in buoy-
ancy periods provides a range over the sorted vertical profile’s stratification for
which a calculation of the model’s effective diffusivity will be performed:
dρre f 1.035g/cm3 4π 2
= − , (18.23)
dz∗ 980cm/sec2 T 2
where T (sec) is the period. With 1 × 60secs < T < 6 × 60 × 60secs defining the
period range, the corresponding vertical density gradient range is
dρre f
10−10 g/cm4 ≤ ≤ 10−5 g/cm4 , (18.24)
dz
204 CHAPTER 18. EFFECTIVE DIANEUTRAL DIFFUSIVITY
dσre f
10−7 g/cm4 ≤ ≤ 10−2 g/cm4 . (18.25)
dz
18.2.6 Negative κe f f
Those advection schemes which contain dispersion, such as centered differenced
advection, have leading order error terms that are not second order, but rather
third order differential operators. Hence, the diagnosis of κe f f for these schemes
18.3. AN EXAMPLE WITH VERTICAL DENSITY GRADIENTS 205
will likely to contain a fair amount of negative values. In turn, negative κe f f may be
interpreted as a sign of dispersion errors, which can create or destroy water masses.
Upon introducing convection into the model, much of these undershoots and over-
shoots created by dispersion are rapidly mixed. In turn, the resulting κe f f should
become positive upon introducing convection.
Another source of negative κe f f apparently can arise simply due to the finite
sampling time and discrete grid, even in the case of pure diffusion. For example,
if there is a mixing event, and if this event is under-sampled in time, it is possible
that the sorted state may have density appear in a non-local manner. Such mixing
events will lead to negative κe f f . The ability to realize such values for κe f f motivates
a sampling time ∆t equal to time step used to evolve the unsorted density.
and the corresponding sorted density state, evolve under the effects of vertical dif-
fusion? Note the grid dimensions for the two states are related through
∆z = 4 ∆z∗ , (18.27)
where z is the vertical coordinate for the unsorted state, and z∗ is the vertical coor-
dinate for the sorted state. For the following, it is convenient to define this state as
that at time (t − ∆t). The potential energies of the unsorted and sorted states are
easily computed to be
E p (t − ∆t) = 56 ρo g ∆z V (18.28)
Ere f (t − ∆t) = 56 ρo g ∆z V (18.29)
E APE (t − ∆t) = 0 (18.30)
where V is the volume of the grid cells, and ρo is the density scale. The zero APE is
due to the absence of horizontal density gradients.
2 2 2 2 4
3∆z 4 12 ∆z* = 3 ∆z
4 4 4 4
6 6 6 6 4
4
6
Figure 18.1: The initial density field for the first example. The number in each
box represents the density, given in units of ρo . The left panel shows the density
ρ( x, z, t − ∆t) in the unsorted fluid state, and the right panel shows the density
ρre f ( z∗ , t − ∆t) in the sorted state. Note that the vertical scale ∆z∗ = ∆z/4 for the
sorted state has been expanded for purposes of display.
F z ( x, z, t) = −κ δ z ρ( x, z, t)
ρ( x, z + ∆z, t) − ρ( x, z, t) (18.32)
≈ −κ .
∆z
F z ( x, z, t) is defined at the top face of the density grid cell whose center has position
( x, z). In the following, it is useful to introduce the dimensionless quantity
This number arises from the chosen discretization of the diffusion equation. For
linear stability of the discretization, δ(v) < 1 must be maintained.
The top panel of Figure 18.2 shows the vertical diffusive flux through the cell
faces at time t − ∆t, and the bottom panel shows the resulting density field ρ( z, z, t +
∆t). Density in the middle row does not change, whereas the upper row density
increases and the lower row density decreases. The potential energy of this state is
This increase in potential energy is a result of the raised center of mass arising from
the vertical diffusive fluxes.
0 0 0 0
2 2 2 2
3∆z
2 2 2 2
0 0 0 0
2+ δ 2+ δ 2+ δ 2+ δ
4 4 4 4 3∆ z
6- δ 6- δ 6- δ 6- δ
Figure 18.2: Top panel: The vertical diffusive flux F z ( x, z, t − ∆t), in units of
ρo κ /∆z, passing through the faces of the unsorted density grid cells. Bottom panel:
The unsorted density field ρ( x, z, t + ∆t), in units of ρo , where the dimensionless
increment δ is given by δ = 2δ(v) = 4 κ ∆t/(∆z)2 . This density field results from
the vertical convergence of the flux F z ( x, z, t − ∆t).
208 CHAPTER 18. EFFECTIVE DIANEUTRAL DIFFUSIVITY
∆z∗
∗ z∗ ∗
κe f f ( z , t − ∆t) = − F ( z , t − ∆t) .
ρre f ( z∗ + ∆z∗ , t − ∆t) − ρre f ( z∗ , t − ∆t)
(18.38)
The units for κe f f ( z∗ , t − ∆t) are (∆z∗ )2 /∆t. Hence, a value for κe f f ( z∗ , t − ∆t) of
2 δ in Figure 18.3 indicates a dimensional value of
(∆z∗ )2
κe f f ( z∗ , t − ∆t) = 2 δ
∆t
4 ∆t (∆z∗ )2 (18.39)
=κ
(∆z)2 ∆t
= κ /4.
This example illustrates a problem with unstratified parts of the sorted profile.
As evident from Figures 18.1 and 18.3, the 12 sorted boxes are actually three larger
homogeneous boxes, and so the calculation should compute fluxes and diffusivi-
ties for these three boxes rather than for the 12 boxes. Figure 18.4 shows such a
combined system, where there are three boxes each of height ∆z comprising the
sorted state. Repeating the previous calculation for this configuration recovers the
expected κe f f = κ on the two interior interfaces. Note that there is no ad hoc setting
to zero certain values of κe f f associated with unstratified portions of the profile.
As a final note, the potential energy of the sorted state at time t + ∆t is
0 0 0
2+ δ
δ 0 0
2+δ
2δ 0 0
2+δ
3δ 0 0
2+δ
4δ -2 2δ
4
4δ 0 0
4
4δ 0 0
4
4δ 0 0
4
4δ -2 2δ
6- δ
3δ 0 0
6- δ
2δ 0 0
6- δ
δ 0 0
6- δ
0 0 0
Figure 18.3: First panel (left): The sorted density field ρre f ( z∗ , t + ∆t), in units of ρo .
∗
Second panel: The vertical diffusive flux F z ( z∗ , t − ∆t), in units of ρo ∆z∗ /(2 ∆t),
passing through the faces of the sorted density grid cells. Third panel: The vertical
density gradient [ρre f ( z∗ + ∆z∗ , t − ∆t) − ρre f ( z∗ , t − ∆t)]/∆z∗ in units of ρo /∆z∗ .
Fourth panel: The effective diffusivity κe f f ( z∗ , t − ∆t) in units of (∆z∗ )2 /∆t.
210 CHAPTER 18. EFFECTIVE DIANEUTRAL DIFFUSIVITY
0 0 0
2 2+ δ
-2 δ κ
3∆z
4 4
-2 δ κ
6 6−δ
0 0 0
Figure 18.4: First panel (far left): The initial density field ρre f ( z∗ , t − ∆t), consisting
of the combination of the three groups of four homogeneous cells. The values are
given in units of ρo . In this recombined arrangement, ∆z∗ = ∆z. Second panel:
The vertical density gradient [ρre f ( z∗ + ∆z∗ , t − ∆t) − ρre f ( z∗ , t − ∆t)]/∆z∗ , in units
of ρo /∆z∗ . Third panel: The density ρre f ( z∗ , t + ∆t) in units of ρo . Fourth panel:
∗
The diffusive flux F z ( z∗ , t − ∆t). Fifth panel (far right): The effective diffusivity
κe f f ( z∗ , t − ∆t).
18.4. AN EXAMPLE WITH VERTICAL AND HORIZONTAL GRADIENTS 211
which is higher than the initial potential energy as a result of the raised center of
mass. The APE remains unchanged
2 4 6 8 6
4 6 8 10 3∆z 6 12 ∆z* = 3 ∆z
6 8 10 12 8
10
10
12
Figure 18.5: The initial density field for the horizontal and vertical diffusion ex-
amples. The number in each box represents the density, given in units of ρo . The
left panel shows the density ρ( x, z, t − ∆t) in the unsorted fluid state, and the right
panel shows the density ρre f ( z∗ , t − ∆t) in the sorted state. Note that the vertical
scale ∆z∗ = ∆z/4 for the sorted state has been expanded for purposes of display.
F z ( x, z, t) = −κ δ z ρ( x, z, t)
ρ( x, z + ∆z, t) − ρ( x, z, t) (18.46)
≈ −κ .
∆z
F z ( x, z, t) is defined at the top face of the density grid cell whose center has height
z. The top panel of Figure 18.6 shows the vertical diffusive flux through these faces
at time t − ∆t, and the bottom panel shows the resulting density field ρ( x, z, t +
18.4. AN EXAMPLE WITH VERTICAL AND HORIZONTAL GRADIENTS 213
∆t). Density in the middle row does not change, whereas the upper row density
increases and the lower row density decreases. The potential energy of this state is
which is higher than the initial potential energy as a result of the raised center of
mass.
0 0 0 0
2 2 2 2
3∆z
2 2 2 2
0 0 0 0
2+ δ 4+ δ 6+ δ 8+ δ
4 6 8 10 3∆ z
6- δ 8- δ 10- δ 12- δ
Figure 18.6: Top panel: The vertical diffusive flux F z ( x, z, t − ∆t), in units of
ρo κ /∆z, passing through the faces of the unsorted density grid cells. Bottom
panel: The unsorted density field ρ( x, z, t + ∆t), in units of ρo , where δ = 2 δ(v) =
4 κ ∆t/(∆z)2 . This is the density field resulting from the vertical convergence
of the flux F z ( x, z, t − ∆t). The potential energy of this field is E p (t + ∆t) =
ρo g ∆z V (110 + 16δ(v) ).
∗
where ρre f ( z∗ , t) is the sorted state’s density. F z ( z∗ , t) is defined at the top face of
the sorted density grid cell whose center has height z∗ . Given the time tendency for
the sorted state, the flux is diagnosed through
∗
z∗ ∗ z∗ ∗ ∗ ∆z
F ( z , t − ∆t) = F ( z − ∆z , t − ∆t) − [ρre f ( z∗ , t + ∆t) − ρre f ( z∗ , t − ∆t)].
2∆t
(18.50)
The left panel of Figure 18.7 shows the sorted density field ρre f ( z , t + ∆t), and∗
∗
the second panel shows the diagnosed vertical diffusive flux F z ( z∗ , t − ∆t). The
third panel shows the vertical density gradient [ρre f ( z∗ + ∆z∗ , t − ∆t) − ρre f ( z∗ , t −
∆t)]/∆z∗ . Note the regions of zero stratification. The fourth panel shows the effec-
tive diffusivity κe f f ( z∗ , t − ∆t), which is diagnosed from the relation
∆z∗
∗ z∗ ∗
κe f f ( z , t − ∆t) = − F ( z , t − ∆t) .
ρre f ( z∗ + ∆z∗ , t − ∆t) − ρre f ( z∗ , t − ∆t)
(18.51)
The units for κe f f ( z∗ , t − ∆t) are (∆z∗ )2 /∆t. In addition, consistent with the dis-
cussion in Section 18.3.2, the effective diffusivity for the interfaces on top of un-
stratified water are multiplied by the number of unstratified boxes. A value for
κe f f ( z∗ , t − ∆t) of δ in Figure 18.7 indicates a dimensional value of
(∆z∗ )2
κe f f ( z∗ , t − ∆t) = δ
∆t
4 ∆t (∆z∗ )2
= κ
(∆z)2 ∆t
= κ /4. (18.52)
which is higher than the initial potential energy as a result of the raised center of
mass. The APE is therefore given by
0 0 0
2+ δ
δ -2 δ /2 δ
4
δ 0 0
4 +δ
2δ -2 δ 3δ
6- δ
δ 0 0
6
δ 0 0
6+ δ
2δ -2 δ 3δ
8- δ
δ 0 0
8
δ 0 0
8+δ
2δ -2 δ 2δ
10 - δ
δ 0 0
10
δ -2 δ /2
12 - δ
0 0 0
Figure 18.7: Left panel: The sorted density field ρre f ( z∗ , t + ∆t), in units of ρo . Sec-
∗
ond panel: The vertical diffusive flux F z ( z∗ , t − ∆t), in units of ρo ∆z∗ /∆t, passing
through the faces of the sorted density grid cells. Third panel: The vertical density
gradient [ρre f ( z∗ + ∆z∗ , t − ∆t) − ρre f ( z∗ , t − ∆t)]/∆z∗ in units of ρo /∆z∗ . Fourth
panel: The effective diffusivity κe f f ( z∗ , t − ∆t) in units of (∆z∗ )2 /∆t. The four κe f f
values which are on top of unstratified portions of the ρre f ( z∗ , t − ∆t) profile have
been multiplied by the number of unstratified boxes which lie directly beneath it.
216 CHAPTER 18. EFFECTIVE DIANEUTRAL DIFFUSIVITY
F x ( x, z, t) = − A δ x ρ( x, z, t)
ρ( x + ∆x, z, t) − ρ( x, z, t) (18.56)
≈ −A .
∆x
F x ( x, z, t) is defined at the east face of the density grid cell whose center has position
( x, z). The top panel of Figure 18.8 shows the horizontal diffusive flux through
these faces at time t − ∆t, and the bottom panel shows the resulting density field
ρ( x, z, t + ∆t). The potential energy of this state is the same as the initial potential
energy, since the horizontal fluxes are parallel to the geopotential
∗
F z ( z∗ , t) = −κe f f ( z∗ , t) δ z∗ ρre f ( z∗ , t)
ρre f ( z∗ + ∆z∗ , t) − ρre f ( z∗ , t) (18.59)
∗
≈ −κe f f ( z , t) ,
∆z∗
∗
where ρre f ( z∗ , t) is the sorted state’s density. F z ( z∗ , t) is defined at the top face of
the sorted density grid cell whose center has height z∗ . Given the time tendency for
the sorted state, the flux is diagnosed through
∆z∗
z∗ ∗ z∗ ∗ ∗
F ( z , t − ∆t) = F ( z − ∆z , t − ∆t) − [ρre f ( z∗ , t + ∆t) − ρre f ( z∗ , t − ∆t)].
2∆t
(18.60)
The left panel of Figure 18.4.2.2 shows the sorted density field ρre f ( z∗ , t + ∆t), and
∗
the second panel shows the diagnosed vertical diffusive flux F z ( z∗ , t − ∆t). The
third panel shows the vertical density gradient [ρre f ( z∗ + ∆z∗ , t − ∆t) − ρre f ( z∗ , t −
∆t)]/∆z∗ . Note the regions of zero stratification. The fourth panel shows the effec-
tive diffusivity κe f f ( z∗ , t − ∆t), which is diagnosed from the relation
∆z∗
∗ z∗ ∗
κe f f ( z , t − ∆t) = − F ( z , t − ∆t) .
ρre f ( z∗ + ∆z∗ , t − ∆t) − ρre f ( z∗ , t − ∆t)
(18.61)
18.4. AN EXAMPLE WITH VERTICAL AND HORIZONTAL GRADIENTS 217
0 δ δ δ 0
3∆z
0 δ δ δ 0
0 δ δ δ 0
2+ δ 4 6 8- δ
4+ δ 6 8 10 - δ 3∆ z
6+ δ 8 10 12 - δ
Figure 18.8: Top panel: The horizontal diffusive flux F x ( x, z, t − ∆t), in units
of ρo A/∆x, passing through the faces of the unsorted density grid cells. Bottom
panel: The unsorted density field ρ( x, z, t + ∆t), in units of ρo , where δ = 2 δ(h) =
4 A ∆t/(∆x)2 . This is the density field resulting from the vertical convergence of the
flux F x ( x, z, t − ∆t). The potential energy of this field is E p (t + ∆t) = 110 ρo g ∆z V.
218 CHAPTER 18. EFFECTIVE DIANEUTRAL DIFFUSIVITY
The units for κe f f ( z∗ , t − ∆t) are (∆z∗ )2 /∆t. For example, a value for κe f f ( z∗ , t − ∆t)
of 3δ /2 in Figure 18.4.2.2 indicates a dimensional value of
(∆z∗ )2
κe f f ( z∗ , t − ∆t) = (3 δ /2)
∆t
∆t (∆z∗ )2
=6A (18.62)
(∆x)2 ∆t
∗ 2
∆z
=6A .
∆x
For the special case of ∆x = ∆z = 4∆z∗ , the effective diffusivity is 3A/8. Note
that if the patch proposed in Section 18.3.3 is used, then the 3δ /2 diffusivity would
become 9δ /2, leading to the possibility for an effective diffusivity of 9 A/8, which
is impossible.
As a final note, the potential energy of the sorted state at time t + ∆t is
which is higher than the initial potential energy as a result of the raised center of
mass. The APE is given by
0 0 0
2+ δ
δ -2 δ /2
4
δ 0 0
4 +δ
2δ -2 δ
6
2δ 0 0
6
2δ 0 0
6+δ
3δ -2 3δ/2
8- δ
2δ 0 0
8
2δ 0 0
8
2δ -2 δ
10 - δ
δ 0 0
10
δ -2 δ/2
12 - δ
0 0 0
Figure 18.9: First panel (far left): The sorted density field ρre f ( z∗ , t + ∆t) in units of
∗
ρo . Second panel: The vertical diffusive flux F z ( z∗ , t − ∆t), in units of ρo ∆z∗ /∆t,
passing through the faces of the sorted density grid cells. Third panel: The vertical
density gradient [ρre f ( z∗ + ∆z∗ , t − ∆t) − ρre f ( z∗ , t − ∆t)]/∆z∗ in units of ρo /∆z∗ .
Fourth panel: The effective diffusivity κe f f ( z∗ , t − ∆t) in units of (∆z∗ )2 /∆t. The
four κe f f values which are on top of unstratified portions of the ρre f ( z∗ , t − ∆t)
profile have been multiplied by the number of unstratified boxes which lie directly
beneath it.
220 CHAPTER 18. EFFECTIVE DIANEUTRAL DIFFUSIVITY
CHAPTER
NINETEEN
Age tracers
Contents
Age tracers are perhaps the simplest nontrivial passive tracers commonly used
in ocean modeling. They provide the modeler with a visual means to deduce the
time that a water parcel has spent circulating after having been in contact with a
certain part of the domain. The purpose of this chapter is to present the basics for
how mom4 computes age tracers.
where dM = ρ dV is the total mass of the parcel of seawater, and C = dMC /dM
is the mass of tracer per mass of seawater. C is often termed the tracer concentration
or tracer mass fraction. Notably, since the seawater mass and tracer mass are both
scalar quantities, the tracer concentration is likewise a scalar field. Hence, trac-
ers are often called scalars in the fluid dynamics literature. Additionally, when the
tracer concentration C is uniform throughout the ocean, the budget for tracer mass
reduces to the budget for seawater mass. This observation, which follows from
the definition of C, provides an important compatibility constraint between mass
and tracer budgets that should be maintained in numerical simulations. Otherwise,
spurious source/sinks can cause tracer or mass to form in unphysical manners.
222 CHAPTER 19. AGE TRACERS
where we used mass conservation d(ρ dV )/dt = 0. Combining the Eulerian form
of material tracer conservation dC /dt = (∂t + v · ∇) C = 0 with mass conservation
ρ,t + ∇ · (ρ v) = 0 leads to the Eulerian conservation law for tracer mass per volume
ρ C within in a seawater parcel
(ρ C ),t + ∇ · (ρ C v) = 0. (19.3)
Additional terms appear on the right hand side when sources or sinks of tracer are
found within the fluid or at the boundaries.
For purposes of establishing the momentum and tracer budgets, the mass of
a Boussinesq seawater parcel is given by dM = ρo dV. Material conservation of
tracer mass (ρo dV ) C still leads to material conservation of tracer concentration:
dC /dt = 0, since now the parcel’s volume is conserved. Combining material con-
servation of tracer concentration with parcel volume conservation, ∇ · v = 0, leads
to the Eulerian conservation law
C,t + ∇ · (v C ) = 0. (19.4)
There are two main ways that the mass of tracer within a seawater parcel can
change. First, there can be tracer-dependent sources or sinks of tracer mass ρ S (C)
within the fluid domain or at the boundaries. Second, there can be irreversible
molecular or turbulent mixing effects that move tracer mass between parcels. Many
of these mixing effects can be mathematically represented by the convergence of
a density weighted tracer flux ρ F. Hence, for a non-Boussinesq fluid, the time
tendency for density weighted tracer concentration is written
S (C) = 1. (19.7)
This prescription specifies that the age field C has units of time. Hence, the age
tracer is somewhat distinct from the material tracers discussed in Section 19.1,
19.2. AGE TRACERS 223
Miscellaneous Topics
227
M ISCELLANEOUS T OPICS
This part of the MOM4 Guide contains discussions of various topics that do not
cleanly fall into other parts.
228
CHAPTER
TWENTY
Contents
The purpose of this chapter is to present certain issues related to the equation
of state used to compute density and its partial derivatives.
20.1 Introduction
Density is needed to compute the hydrostatic pressure. Additionally, it determines
the static stability of a vertical fluid column. Finally, its derivatives with respect to
potential temperature and salinity
∂ρ
ρ,θ = (20.1)
∂θ p,s
∂ρ
ρ,s = (20.2)
∂s p,θ
set the neutral directions used for sub-grid-scale tracer transport. Therefore, it is im-
portant that the equation of state be accurate over the range of temperature, salinity,
and pressure values occurring in ocean simulations.
In early versions of MOM, density was computed according to the Bryan and
Cox (1972) cubic polynomial approximation to the UNESCO equation of state (Gill
(1982)). That approach was quite useful for certain problems. Unfortunately, it
has limitations that are no longer acceptable for global climate modeling. First,
the polynomials are fit at discrete depth levels. The use of partial cells makes this
approach cumbersome since with partial cells it is necessary to generally compute
density at arbitrary depths. Second, the cubic approximation is inaccurate for many
230 CHAPTER 20. EQUATION OF STATE CONSIDERATIONS
regimes of ocean climate modeling, such as wide ranges in salinity associated with
rivers and sea ice. For these two reasons, a more accurate method is desired.
Two equations of state (EOS) are currently available in MOM4 for computing
density. The first is a linear equation of state whereby density is a linear func-
tion only of potential temperature. This EOS is relevant only for idealized simu-
lations with the Boussinesq approximation. The second EOS is that proposed by
McDougall et al. (2003). As they argue, their EOS is more accurate than the UN-
ESCO EOS due to the use of more accurate empirical data as reported in Feistel
(1993) and Feistel and Hagen (1995). Even given its high degree of accuracy, this
EOS has been found to be quite efficient.
• The McDougall et al. (2003) EOS is very efficient, so its cost is much less than
earlier experiences coding the UNESCO form taken from Gill (1982) or Jackett
and McDougall (1995).
• Use of a realistic EOS in a z-coordinate model comes without the price of com-
plex algorithms, in contrast to the situation with isopycnal models. Hence, it
makes sense to provide z-models with the capability of computing highly ac-
curate equilibrium thermodynamic properties.
with
Due to the absence of pressure effects, this density is most correctly thought of as a
potential density. Hence, without compressibility, its use for computing pressure is
self-consistent only when employing the Boussinesq approximation.
The EOS has 25 terms. Its general form is motivated by that of Wright (1997).
Appendix B to McDougall et al. (2003) provides the following equation for in situ
density ρ written in terms of pressure, salinity, and potential temperature
P1 (s, θ, p)
ρ(s, θ, p) = , (20.11)
P2 (s, θ, p)
A check value for this equation is ρ = 1033.213387 kg. m−3 with s = 35psu, θ =
20◦ C, and p = 2000db = 2 × 107 Pa. The polynomial functions P1 and P2 are given
by
P1 = ao + a1 θ + a2 θ 2 + a3 θ 3 + a4 s + a5 s θ + a6 s2
+ a7 p + a8 p θ 2 + a9 p s + a10 p2 + a11 p2 θ 2 (20.13)
P2 = bo + b1 θ + b2 θ 2 + b3 θ 3 + b4 θ 4 + b5 s + b6 s θ + b7 s θ 3 + b8 s3/2 + b9 s3/2 θ 2
+ b10 p + b11 p2 θ 3 + b12 p3 θ. (20.14)
232 CHAPTER 20. EQUATION OF STATE CONSIDERATIONS
P1 = ao + θ ( a1 + θ ( a2 + a3 θ )) + s ( a4 + a5 θ + a6 s)
+ p ( a7 + a8 θ 2 + a9 s + p ( a10 + a11 θ 2 )) (20.15)
P2 = bo + θ (b1 + θ (b2 + θ (b3 + θ b4 ))) + s (b5 + θ (b6 + b7 θ 2 ) + s1/2 (b8 + b9 θ 2 ))
+ p (b10 + p θ (b11 θ 2 + b12 p)). (20.16)
!
∂ρ ∂P1 ∂P2
1 1
=ρ − (20.17)
∂θ s,p P1 ∂θ s,p P2 ∂θ s,p
!
∂ρ ∂P1 ∂P2
1 1
=ρ − . (20.18)
∂s θ,p P1 ∂s θ,p P2 ∂s θ,p
" #
∂ρ ∂P1 ∂P2
= ( P2 )−1 −ρ (20.19)
∂θ s,p ∂θ s,p ∂θ s,p
" #
∂ρ ∂P1 ∂P2
−1
= ( P2 ) −ρ (20.20)
∂s θ,p ∂s θ,p ∂s θ,p
where ( P2 )−1 can be saved in a temporary array, thus reducing the number of divi-
sions from four to one.1 The partial derivatives are given by
∂P1
= a1 + 2 a2 θ + 3 a3 θ 2 + a5 s + 2 a8 p θ + 2 a11 p2 θ (20.21)
∂θ s,p
∂P1
= a4 + a5 θ + 2 a6 s + a9 p (20.22)
∂s θ,p
∂P2
= b1 + 2 b2 θ + 3 b3 θ 2 + 4 b4 θ 3 + b6 s + 3 b7 s θ 2 + 2 b9 s3/2 θ (20.23)
∂θ s,p
+ 3 b11 p2 θ 2 + b12 p3
∂P2
= b5 + b6 θ + b7 θ 3 + (3 / 2 ) b8 s1/2 + (3 / 2 ) b9 s1/2 θ 2 (20.24)
∂s θ,p
∂P1
= a1 + θ (2 a2 + 3 a3 θ ) + a5 s + 2 p θ ( a8 + a11 p) (20.25)
∂θ s,p
∂P1
= a4 + a5 θ + 2 a6 s + a9 p (20.26)
∂s θ,p
∂P2
= b1 + θ (2 b2 + θ (3 b3 + 4 b4 θ )) + s (b6 + θ (3 b7 θ + 2 b9 s1/2 )) (20.27)
∂θ s,p
+ p2 (3 b11 θ 2 + b12 p)
∂P2
= b5 + θ ( b6 + b7 θ 2 ) + ( 3 / 2 ) s1/2 ( b8 + b9 θ 2 ) . (20.28)
∂s θ,p
234 CHAPTER 20. EQUATION OF STATE CONSIDERATIONS
CHAPTER
TWENTYONE
Contents
The purpose of this chapter is to present the method used in mom4 for pre-
scribing open boundary conditions. This chapter was written by Martin Schmidt
([email protected]) and it is under revision.
21.1 Introduction
Numerical circulation models of marginal seas with biological, chemical and sed-
iment dynamic components require a high model resolution and involve a large
number of variables. Hence, a regional model is the only realistic implementation
236 CHAPTER 21. OPEN BOUNDARY CONDITIONS
laws but requires many assumptions and approximations. Differently from the
other model components it should be considered as an engineering tool, which
must be tuned to find a reasonable compromise for the special model purposes.
The implementation of OBC ihas been subject of many papers and technical re-
ports. The implementation depends strongly on the model grid, the present scheme
is for the Arakaw B-grid and follows in general the approach of Stevens (1991). It
has been tested successfully in the FRAM model (Stevens, 1991) and in the com-
munity modeling effort (CME) , (e.g., Döscher and Redler, 1995). They have also
been used in a GFDL-model of the North Atlantic in the framework of MAST2-
DYNAMO (Dynamo, 1994) as well as in a regional model of the subpolar North
Atlantic (Redler and Böning, 1996). These experiments were based on Redlers im-
plementation of OBC in MOM-1.
ROMS others ...
For the explicit free surface code used in MOM-3 the implementation of open
boundary conditions was modified. Especially the calculation of the phase speed
of barotropic waves must be refined. The modifications of the free surface open
boundary conditions in MOM-3 are based on the experience with the methods
developed by A. Mutzke (Mutzke, 1996) for a circulation model of the Baltic Sea
(Seifert, Schmidt, Fennel, 1999). driven by 3 hourly gauges in the Skagerak.
∂η
U̇ + gH = X, (21.1)
∂x
∂η
V̇ + gH = Y (21.2)
∂y
∂U ∂V
η̇ = Q − + . (21.3)
∂x ∂y
η denotes the sea level elevation, X and Y are the wind stress components, Q is
a fresh water flux, H the total depth and g the gravitational accelleration. These
equations can be combined to a single equation for the sea level elevation,
∂ ∂η ∂ ∂η ∂X ∂Y
− η̈ + gH + gH = + − Q̇. (21.4)
∂x ∂x ∂y ∂y ∂x ∂y
We consider an initial value problem for time t = 0. The sea surface has a well
238 CHAPTER 21. OPEN BOUNDARY CONDITIONS
2 ∂
2
2
ω +c G ( x, x0 , ω) = δ ( x − x0 ). (21.14)
∂x02
The boundary conditions at the edge of the model area will be specified later.
Subtracting equation (21.14) multiplied with G from equation (21.11) multiplied
with η, integrating over x0 the general solution reads
Z xr
η( xω) = dx0 G ( x, x0 , ω) F ( x0ω)+
xl
x 0 = xr
∂G ( x, x0 , ω) ∂η( x0 )
2 0 0
c η( x ) − G ( x, x , ω ) (21.15)
∂x0 ∂x0 x0 = xl
21.2. ONE-DIMENSIONAL, LINEAR, NON-ROTATING WAVES 239
xl and xr are the left and the right boundary of the channel. Hence, specifying
appropriate boundary conditions for the Green function at x0 = xl and x0 = xr may
help to simplify the problem considerably.
To discuss the open boundary problem we split the channel at x = a and ask
only for the sea level within an area of interest, say at x > a. The point x = a is
the open boundary. The sea level elevation at all points x > a consists now of two
contributions, ηl describes waves propagating eastward from the area left of the
boundary and ηr which stands for all waves generated at all points x > a within
the area of interest. Also the forcing function F can be devided formally in two
parts,
F ( xω) = θ ( x − a) Fr ( xω) + θ ( a − x) Fl ( xω) (21.20)
Fr is the forcing in the area east of x = a, Fl decribes all forcing west of x = a. ηl
reads
Z a
1 x− x0
ηl ( xω) = dx0 eiω c Fl ( x0ω),
2iωc −∞
i ωc ( x− a)
= ηl ( aω)e . (21.21)
the result for ηr is
Z ∞
1 | x− x0 |
ηr ( xω) = dx0 eiω c Fr ( x0ω).
2iωc a
(21.22)
The inverse Fourier transforms are
a 0
dω 1 t− x−c x
Z Z
−iω
ηl ( xt) = dx0 e Fl ( x0ω)
2π 2iωc −∞
x−a
= ηl a, t − (21.23)
c
240 CHAPTER 21. OPEN BOUNDARY CONDITIONS
with
x 0
dω 1 −iω t− x−c x
Z Z
ηr< ( xt)= dx0 e Fr ( x0ω)
2π 2iωc a
Z ∞
x− x0
dω 1
Z
0 −iω t+ c
>
ηr ( xt) = dx e Fr ( x0ω) (21.24)
2π 2iωc x
Hence, considering the sea level elevation η = ηl + ηr at a certain point x, it is the
superposition of three contributions. ηl depends on x − ct and decribes eastward
travelling waves generated somewhere at locations x < a. The first contribution of
ηr , ηr< ( x − ct) depending on x − ct is a superposition of waves generated at loca-
tions between a and x travelling eastward, the second contribution of ηr , ηr> ( x + ct),
describes westward moving waves coming from locations eastward of x. From a
mathematical point of view the problem is solved, if the Green function and all
forces for x > a are known, and a time series of incoming waves at x = a is pre-
scribed. However, the separation of the solution in incoming and outgoing waves
is not part of a numerical scheme.
0 ∂G ( xx , ω ) ∂η( x0 )
0
2 0
η( xω) = ηr ( xω) − c η( x ) − G ( xx , ω) . (21.25)
∂x0 ∂x0 x0 = a
With
∂G ( xx0 , ω) 1 ω
c2 = − ei c ( x− a) (21.27)
∂x0 x0 = a 2
∂G ( xx0 , ω) 1 i ω ( x− a)
c2 = e c (21.28)
∂x x0 = a 2
∂G 2 ( xx0 , ω) iω ω
c2 = − ei c ( x− a) (21.29)
∂x∂x0 x0 = a 2c
(21.30)
since η( aω) is influenced also from the area at x < a. However, with the help of
the above relations equation (21.25) can be rewritten where either the sea level or
alternatively the sea level gradient at x = a must be prescribed.
ω
η( xω) = ηr ( xω) + ei c (x−a) (η( aω) − ηr ( aω)) (21.33)
i ωc ( x− a)
= ηr ( xω) + e ηl ( aω) (21.34)
This is the same form as equation (21.21) but ηl ( aω) is now an unknown boundary
value which must be prescribed. Alternatively, the gradient of ηl can be prescribed,
η0 ( x) = η0θ ( x − x0 ), (21.36)
where the sea surface has a level η0 east of x = x0 and 0 elsewhere. To position
of the initial step is assumed within the model domain, x0 > a. Hence, at least for
small times ηl = 0 should be valid.
i η0 ω
ω
ηr ( xω) = θ ( x − x0 ) 2 − ei c (x−x0 ) + θ ( x0 − x) ei c (x0 −x) (21.37)
2ω
solution follows
c ∂ηr ( xω)
ηr ( aω) = − | x= a (21.38)
iω ∂x
c ∂ηl ( xω)
ηl ( aω) = | x= a . (21.39)
iω ∂x
This is equivalent to
It means, that ηr at the boundary x = a consists of westward going waves only but
ηl is spreading only eastward. Hence, requiring for the total solution η
at x > a. A similar boundary condition was originally derived for the theory of
electromagnetic waves to remove incoming waves from the far field solution for an
oscillating dipole antenna and is called commonly Sommerfeld radiation condition
(see e.g. Sommerfeld, 1966).
In the context of our Greens function approach, the equations (21.40) and (21.40)
have been derived and reflect the boundary conditions applied for the Greens func-
tion, which means, that all relevant sources of waves must have a finite distance
from the origin. Hence, here these equations are not a boundary condition but a
general property of the wave field. On the other hand equation (21.42) restricts
the manifold of solutions of the wave equation. It corresponds to ηl ( a) = 0 and is
suitable as boundary condition which filters out incoming waves.
All deviations between the prescribed and the internally forced sealevel generate
waves propagating into the model domain. Only if the prescribed sealevel is the
exact solution no inwards propagating waves occur. Hence, boundary values and
forcing must be consistent data sets to avoid artificial values for ηl ( a) and artificial
waves by reflection of the internal solution at the model boundary. This is an un-
realistic constraint for oceanographic applications where boundary data are taken
in most cases from other models or measurements. Surface gravity waves are vary-
ing at timescales of seconds to minutes but boundary values are mostly availabe
with timescales of hours, days or even month. An open boundary with prescribed
slowly varying external values would simply reflext outgoing waves back into the
model area.
From this point of view the problem to find the sealevel within a given model
domain from the equations, initial values, the forcing and prescribed boundary
values is an ill-posed problem which can be solved only approximately. As a guide
how this can be accomplished, the special case of no incoming waves can be used.
In this case the Sommerfeld radiation condition applies to the total solution and no
boundary values are needed. Hence, time tendency of the total η at the boundary
x = a consists of an contribution governed by a Sommerfeld radiation condition
21.3. THE ROTATING CASE 243
and an external contribution. The latter is only present, when the internal time
tendency leads to an result different from the prescribed extrenal value,
∂η( xt) ∂η( xt) ηext (t) − η( at)
| x= a = c | x= a + . (21.45)
∂t ∂x τ
This equation can be read as follows. If the solution follows the externally pre-
scribed value ηext , there are no externally driven incoming waves. If there is a
difference, the solution at the boundary is slowly relaxed to the prescribed value,
the parameter τ rules the relaxation time. In more refined schemes (Orlanski ...) pa-
rameterisations like τ = τ (c) is used to relax only, if incoming waves are detected.
i i i i i i
B+1 B B-1 B-1 B B+1 velocity index
i i i i i i
B+1 B B-1 B-1 B B+1 tracer index
Figure 21.1: The geometry of western (left) and eastern (right) open boundaries.
Note the index convention explained below.
j
B-1
j
B-1
j
B
j
B
j
B+1
j
B+1
velocity index tracer index
Figure 21.2: The geometry of southern (lower) and northern (upper) open bound-
aries. Note the index convention explained below.
21.5. MOMENTUM EQUATIONS 245
Here qk and q⊥ are the spatial coordinates parallel or perpendicularly to the bound-
ary, c is the wave phase speed and v ad the advection velocity. Assuming v ad to be
constant T obeys the differential equation
∂T ∂T
= ± (c + v ad ) . (21.59)
∂t ∂q⊥
The negative sign corresponds to waves traveling in the direction of q⊥ , the positive
sign corresponds to waves traveling in the opposite direction. Thus, choosing either
the positive or the negative sign waves entering the model area from outside can
be removed from the solution keeping only outgoing waves. This so called Som-
merfeld radiation condition (see e.g. Sommerfeld, 1966) was originally derived for
the theory of electromagnetic waves to remove incoming waves from the far field
solution for an oscillating dipole antenna.
Hence, the modified tracer equation at an open boundary point is the superpo-
sition of the radiation condition and the diffusion term and reads
∂T c T + v ad ∂T ∂T ∂T
+ = FT + ST , cT = −a / (21.60)
∂t a ∂φ ∂t ∂φ
∂T c T + u ad ∂T ∂T ∂T
+ = FT + ST , c T = − a cos φ / (21.61)
∂t a cos φ ∂λ ∂t ∂λ
−1
TBn−1 − TBn− 1 q⊥,B−1 − q⊥,B−2
cnnum = − n n , (21.62)
TB−1 − TB−2 tn − tn−1
where the superscript ”n“ denotes the time step. The CFL-criterion provides a limit
cmax for the phase speed,
q⊥,B−1 − q⊥,B−2
cmax = . (21.63)
2∆t
21.7. THE SEA LEVEL 247
It may happen that the phase speed is directed into the model domain. In this case
it is set to zero.
For outgoing waves, the gradient of T in the advective term is calculated from TB
and TB−1
The diffusion operator F T requires the knowledge of TB+1 . For a passive bound-
ary the approximation
TB+1 = TB (21.67)
are possible.
This ”passive“ boundary condition may not be sufficient for many purposes.
After running a ”passive“ boundary over a long period, the tracer near the bound-
ary will be determined completely by processes in the model domain. As an ex-
ample consider a marginal sea with a strong fresh water surplus. There will be an
estuarine circulation with a more or less permanent outflow of brackish water in
a surface layer and inflow near the bottom. However, the salinity of the inflow-
ing water will be reduced too after some time by vertical mixing processes and the
model results will suffer from underestimated stratification.
To overcome this problem information on the tracer concentration in the adja-
cent sea must be provided for the model. The simple approximation
TB+1 = T0 , (21.69)
where T0 may stem from an appropriate database, improves the performance of the
diffusion operator which in turn may invoke wave like processes traveling from the
boundary into the model. Using an upstream formulation for the tracer gradient
in the advective term, this can switch on an inflow through the open boundary.
However, waves of a small amplitude but with a high phase speed may disturb
this scheme. Thus, the source term S T can be used for a controlled restoring to
prescribed boundary values.
and depends only on the argument α = ~k ·~r − ωt where ~k is the wave number and
ω is the wave frequency. The waves move with phase velocity
ω ~k
~c = . (21.71)
k k
From the derivatives (in Cartesian co-ordinates)
ηt = −ωηα
η x = k x ηα
η y = k y ηα (21.72)
ηt = −c x η x − c y η y . (21.73)
with
ω2
c2 = ,
k2
ω ηt
ax = =− ,
kx ηx
ω ηt
ay = =− ,
ky ηy
c2 ηt η x
cx = − =− 2 ,
ax η y + η2x
c2 ηt η y
cy = − =− 2 ,
ay η y + η2x
(21.74)
Notable, c2 is obeys
c−2 = a− 2 −2
x + ay , (21.75)
c2 = c2x + c2y , (21.76)
(21.77)
For free, barotropic, linear surface waves without Coriolis force, the velocity
components depend also only on α and an equation like equation (21.73) can be
derived also directly from the sea level equation
η t = − U x − Vy + q w , (21.78)
ηt = −c x η x − c y η y + qw , (21.79)
21.7. THE SEA LEVEL 249
The more general form with the additional contribution from the fresh water flux,
qw , could be found also from equation (21.70) considering the quantity η̃ = η(~k ·
~r − ωt) + qw t.
So far, in its differential form equation(21.79) is an identity. To transform it into a
prognostic equation to be used at the boundary, the phase speed must be diagnosed
near the model boundary. Icoming waves are suppressed by the radiation condition
~c · ~n > 0, (21.80)
where ~n is a vector directed outward of the model domain normally to the bound-
ary and ~c has the components (c x , c y ).
Considering a western boundary, the condition c x < 0 permits only outgoing
waves, at an eastern boundary c x > 0 is required. In the same manner, at a northern
boundary c y > 0 and at a southern boundary c y < 0 applies.
In geophysical co-ordinates, equation(21.79) reads
∂ηt 1 ∂η ∂η
1
= − + + qw , (21.81)
∂t a cφ ∂φ a cλ cos φ ∂λ
∂η 1 ∂η
∂t a ∂φ
cφ = − 2 2 (21.82)
1 ∂η ∂η
1
a ∂φ + a cos φ ∂φ
∂η 1 ∂η
∂t a cos φ ∂λ
cλ = − 2 2 . (21.83)
1 ∂η 1 ∂η
a ∂φ + a cos φ ∂φ
(21.84)
As for the tracers, the numerical approximation cnnum for the phase speed at time
step n is calculated in the model interior,
ηnB−1 − ηnB− 1
−1 q⊥,B−1 − q⊥,B−2
cnnum = − , (21.93)
ηnB−1 − ηnB−2 tn − tn−1
It is assumed, that the time step is small enough for the CFL-criterion of barotropic
waves.
The calculation of the phases for η appears to more complex than for tracers.
The phase speed is very noisy and needs smoothing. This is done by accumulation
with
cnew
η = cmax , |cnnum | ≥ |cnmax | , outgoing waves (21.96)
cnew
η = cnnum , |cnnum | < |cnmax | , outgoing waves (21.97)
cnew
η = 0, incoming waves (21.98)
For a very small sea level gradient, the phase cannot be calculated. In this case the
approximation
is used.
the model is selfreferencing and small errors in the open boundary conditions may
accumulte and cause large errors in the mass and tracer bilance of the model.
Hence, an appropriate relaxation procedure is needed to correct this model
shortcoming. However, the sea level profile along the open boundary prescribes
the mass transport trough the boundary. Near the coast the sea level profile is
mainly adjusted by Kelvin waves traveling along the coast and leaving behind the
wave fronts a geostrophically adjusted alongshore corrent and the corresponding
sea level profile. As such the relaxation should not disturb the sea level profile at
the boundary but this profile should be generated by the model itself.
As such the relaxation is not done in the sea level equation but in the pressure
gradient term of the barotropic momentum equation at the open boundary. The sea
level at the boundary itself is adjusted only by the Orlanki condition.
To achieve this, the pressure gradients in the equation for the momentum per-
pendicularly to the boundary is modified according
1 1
( p s ) q⊥ → ( p s ) q⊥ + ( ηr ) q⊥ (21.100)
ρo ρo
with
n−1
ηrel − ηnB−
−1
1
( ηr ) q⊥ = aη . (21.101)
∆qbot
More sophisticated schemes can be designed but will be nevertheless rough ap-
proximations which neglect important aspects of the dynamics of the flow across
the boundary.
n−1 φ − φH φ − φM
ηrel (φ) = ηM − ηH (21.103)
φM − φH φM − φH
However, if the model interiour is not commensurable with the prescribed bound-
ary conditions large artificial currents may occur near the boundary.
252 CHAPTER 21. OPEN BOUNDARY CONDITIONS
CHAPTER
TWENTYTWO
Contents
The purpose of this chapter is to describe the formulation of lunar and solar
tidal forcing implemented in mom4. This chapter was written by Harper Simmons
([email protected])
where G is the vertically integrated forcing arising from baroclinic effects, ps is the
pressure associated with undulations of the surface height, p a is the atmospheric
pressure, H is the depth of the ocean, and η is the surface height deviation from a
resting state with z = 0.
Tidal forcing arising from the eight primary constituents (M2, S2, N2, K2, K1,
O1, P1, Q1) (see Gill (1982)) have been added to the forcing for U. The formula-
tion follows Marchuk and Kagan (1989), by considering a tide generating potential
(gηeq ) with corrections due to both the earth tide (1 + k − h) and self-attraction and
loading (α). In this approach, we have
∇ ( ps + p a ) → ∇( ps + p a ) + g ∇ (1 − α )η − (1 + k − h)ηeq . (22.2)
The equilibrium tide, ηeq , arises from the astronomically derived gravity producing
forces. It is modified by several factors. The Love numbers, k and h, named for
the physicist A.L. Love, account for the reduction of the ocean tide because of the
deformation of the solid earth by tidal forces. The Love numbers are frequency
dependent, with 1 + k − h generally close to 0.7 (Wahr (1998)).
254 CHAPTER 22. TIDAL FORCING FROM THE MOON AND SUN
The term α in equation (22.2) accounts for a modification of the ocean’s tidal
response as a result of self-attraction and loading (SAL) (Hendershott (1972)). Self
attraction is the modification of the tidal potential as a result of the redistribution of
the earth and ocean due to the equilibrium tidal forcing. Loading refers to the de-
pression of the earth’s crust by the mounding of tides. Calculation of the SAL term
requires an extremely cumbersome integration over the earth surface, rendering
equation (22.2) an integro-differential equation (Ray (1998)).
Instead of solving the integro-differential form of equation (22.2), MOM4 uses
the scalar approximation to SAL. We feel this is justified since our purpose in in-
troducing tidal forcing is to study the effects of tides on the general circulation, not
the details of the tides themselves. The conjecture is that precise calculation of the
SAL term is not needed for to understand tidal effects on the general circulation.
For the scalar approximation, α is usually set between 0.940 − 0.953. MOM4 uses
α = 0.948. Limitations of the scalar approximation to SAL are discussed by Ray
(1998), who concluded that the scalar approximation introduces phase errors of up
to 30◦ and amplitude errors of 10% into a global scale tidal simulation.
where φ is latitude and λ is longitude. Recognizing that equation (22.3) and (22.4)
require the evaluation of trigonometric functions at every grid point and every
time-step, tidal forcing is introduced into MOM4 in the following mathematically
equivalent form. Making use of the identity
Table 22.1: Frequencies, Love numbers, and amplitude functions for the eight prin-
ciple constituents of tidal forcing available in MOM4.
256 CHAPTER 22. TIDAL FORCING FROM THE MOON AND SUN
CHAPTER
TWENTYTHREE
Contents
The purpose of this chapter is to present a method for parameterizing the inter-
actions between unresolved mesoscale eddies and topography.
23.1 Introduction
Based on statistical mechanics arguments, Holloway (1992) proposed that inter-
actions between mesoscale eddies and topography result in a stress on the ocean
with two important consequences: (1) the ocean is not driven toward a state of rest,
(2) the resulting motion may have scales much larger than the eddy scales. That
is, eddy-topography interactions can generate coherent mean flows on the scale of
the topography. When Holloway described coastal currents that persistently flow
against both the wind forcing and pressure gradient, the response was that it must
be due to King Neptune. Who else? Hence, the effect is referred to colloquially as
the Neptune effect.
The magnitude of the associated topographic stress is dependent on the cor-
relation between pressure p and topographic gradients ∇ H, and this correlation
is largely unknown. But even if it is no larger than 0.1, the resulting topographic
stress could be comparable in magnitude to that from the surface wind.
ψnep = − f L H
(23.1)
unep = ẑ ∧ ∇ψnep
Contents
The purpose of this chapter is to present the two methods used in MOM4 for
temporally discretizing the Coriolis force.
(d/dt + f ẑ ∧ ) u = 0, (24.1)
Here, d/dt is the material time derivative relevant for Lagrangian observers. Mo-
tions which satisfy this equation are termed inertial oscillations and they have period
given by
2π 11.97
Tinertial = = hour (24.3)
f sin φ
where Ω = 7.292 × 10−5 s−1 is the earth’s angular speed. The period of inertial
oscillations is smallest at the North pole where φ = π /2 and Tsmallest ≈ 12 hour.
An explicit temporal discretization of the inertial oscillation equation (24.1) will
be unstable if the time step is longer than some fraction of the inertial period, where
260 CHAPTER 24. TEMPORAL TREATMENT OF THE CORIOLIS FORCE
the fraction depends on details of the time stepping. Coarse resolution models
(models with resolutions on the order of 4-5 degrees) may find this time step con-
straint is the most stringent of the model’s baroclinic momentum processes. To get
around this limitation, a semi-implicit temporal treatment has been traditionally
considered, as in Bryan (1969).
Additional issues with coupling to sea ice may warrant an implicit treatment
even for ocean models run with a momentum time step that well resolves the iner-
tial period. In these cases, temporal details of ocean-ice coupling have been found
to cause enhanced energy at the inertial period. Semi-implicit time stepping of the
Coriolis force may assist in damping this energy.
It is for these reasons that MOM4 provides an option to time step the Coriolis
force either explicitly or semi-implicitly in the baroclinic portion of the model. The
namelist parameter acor sets the level of implicitness, as described in Section 24.4.2.
w = u+iv (24.5)
√
where i = −1 and w should not be confused with the vertical velocity component.
In terms of w, equation (24.4) takes the form
∂ t w = −i f w (24.6)
w = w o ei f t (24.7)
with period
Tinertial = 2 π f −1 . (24.8)
Time discretizing equation (24.6) with a centered leap-frog scheme leads to
with
λ = 2 f ∆τ (24.10)
a dimensionless number. We can write the finite difference solution in terms of an
amplification factor
w(τ + ∆τ ) = G w(τ ). (24.11)
Substituting this ansatz into equation (24.9) leads to the quadratic equation
G2 + i λ G − 1 = 0 (24.12)
24.3. SEMI-IMPLICIT TEMPORAL DISCRETIZATION 261
whose solution is √
−i λ ± −λ 2 + 4
G= . (24.13)
2
If
λ /2 = f ∆τ < 1, (24.14)
then | G | = 1, which means the two finite difference solutions are neutral and stable.
One root is an unphysical mode, known as the leap-frog computational mode, and the
other corresponds to the physical solution. If λ > 2 then | G | > 1 which means both
roots are unstable. Hence, stability requires a time step satisfying
∆τ < f −1 . (24.15)
That is,
Tinertial 2π
= > 2 π, (24.16)
∆τ f ∆τ
meaning the leap-frog scheme remains stable if there are at least 2 π time steps per
inertial period. At the North Pole, this constraint means
For the baroclinic part of the model algorithm, ∆τ < 1.9 hours can be the limit-
ing time step for coarse resolution global models, thus motivating an alternative
approach discussed in Section 24.3.
1 − i λ (1 − γ )
G= . (24.19)
1+iλγ
[1 − γ λ 2 (1 − γ )]2 + λ 2
| G |2 = . (24.20)
[1 + (γ λ )2 ]2
For γ = 0, | G | > 1 which leads to an unstable scheme. For γ = 1/2, | G | = 1
and so the scheme is neutral. With 1/2 < γ ≤ 1, | G | < 1, and so the scheme is
262 CHAPTER 24. TEMPORAL TREATMENT OF THE CORIOLIS FORCE
unconditionally stable. Hence, we arrive at the stability range for the semi-implicit
parameter
1/2 ≤ γ ≤ 1, (24.21)
with γ = 1 yielding the most stable scheme.
As the vertically integrated part of the model algorithm typically uses a time step
much smaller than f −1 , MOM4 discretizes the Coriolis force explicitly when time
stepping the vertically integrated equations. That is, acor is relevant only for the
baroclinic equations.
v · f ẑ ∧ u = 0. (24.35)
This property is respected on the B-grid when we discretize the Coriolis force ex-
plicitly in time
v(τ ) · f ẑ ∧ u(τ ) = 0. (24.36)
However, the semi-implicit treatment does not respect this property since in general
the product
v(τ ) · f ẑ ∧ [(1 − γ ) u(τ − ∆τ ) + γ u(τ + ∆τ )] (24.37)
does not vanish. The only time it will vanish is when the flow is in a steady state.
This limitation of the semi-implicit approach is to be expected, since implicit time
discretizations generally do not conserve energy.
264 CHAPTER 24. TEMPORAL TREATMENT OF THE CORIOLIS FORCE
CHAPTER
TWENTYFIVE
Contents
25.1 Introduction
As discussed by Killworth et al. (1991), discretization of gravity waves on a B-grid
can admit a stationary grid scale checkerboard pattern (see also Messinger (1973)
and Janjić (1974) for atmospheric discussions). This pattern is associated with an
unsuppressed grid splitting that can be initiated through grid scale forcing, such
as topography. That is, the discrete equations admit a checkerboard null mode.
Beckers (1999) also discusses subtleties associated with traditional linear stability.
In our experience, when the model goes unstable due to CFL violations, this grid
mode is the spatial pattern associated with the instability.
Strictly, the grid mode is present only when the surface pressure gradient takes
the form ∇h ps = ρ g ∇h η, where ρ is an averaged surface density. For example,
in the implementation of Killworth et al. (1991), ∇h ps = ρo g ∇h η, and so the null
mode is always present. Tests with the alternative expression ps = ρk=1 g η used in
MOM, in which surface pressure gradients arise from gradients in both the surface
266 CHAPTER 25. CHECKERBOARD NULL MODE
density ρk=1 and surface height, suggest that the null mode is slightly suppressed.
A more significant means to suppress the mode, however, is provided through the
use of time averaging over the barotropic time steps, as well as with nonlinearity
in the shallow water system present when the undulating surface height is fully
incorporated to the dynamics.
Hence, for many purposes, the grid mode is absent or mild. Nonetheless, some
experiments do experience a nontrivial checkerboard pattern in the surface height,
especially in enclosed seas such as the Mediterranean Sea and Hudson Bay. In this
case, the checkerboard mode is continually excited by surface forcing and topogra-
phy, yet there is not enough grid resolution to allow the noise to radiate outward or
as boundary waves. The checkerboard pattern can also be initiated by fresh water
forcing. Although the typical forcing from the atmosphere is large scale, there will
always be some projection onto the grid scale, thus pumping energy into the null
mode. In this case, we have seen the checkerboard pattern appear even over large
basins.
For those cases where the noise is unacceptable, we have considered various
means to suppress it. The following relates our experience provides suggestions
and caveats. We remain incompletely satisfied with the following methods, and so
welcome other suggestions.
press the surface height noise, Killworth et al. (1991) proposed the addition of the
so-called “del-plus / del-cross” operator to the prognostic equation for the free sur-
face height. Their approach is an extension of that suggested by Janjić (1974). The
operator is basically the difference between Laplacians computed in two different
manners, and it preferentially removes the checkerboard pattern by coupling the
grids. This coupling occurs in a scale-selective manner so that the larger scales are
only modestly affected. The construction of this operator requires the definition of
time dependent image points over land to allow for it to conserve volume.
available grid scale filtering. Tests in MOM3 with the del plus / del minus filter,
using the weighting factor suggested by Killworth et al. (1991), transformed large
amount of noise to the vertical velocity field near the surface. Again, the reason
is that this filter is not invisible to the discretized gradients, and so while it acts to
smooth the surface height, it correspondingly adds noise to the surface pressure
gradients which then affect the barotropic velocity.
One attempt to suppress the transfer of noise was to also filter ∇h · U in order
to counteract the noisy pressure gradients. Unfortunately, this approach is prob-
lematic, since doing so breaks the precise connection between the barotropic and
baroclinic systems. As a result, spatially filtering ∇h · U produces a nontrivial spuri-
ous vertical velocity throughout the ocean, and in particular it creates unacceptable
spurious advective fluxes through the ocean bottom.
25.2.2 Suggestions
As mentioned at the beginning of this section, we find the need for filtering to be
greatly reduced due to the use of a nonlinear free surface in which the barotropic
time steps are time averaged. Nevertheless, for those cases in which filtering is
needed, such as for inland seas and when using fresh water forcing, we have found
it sufficient to add a straightforward Laplacian to the free surface height equation
with a reasonably small diffusivity. The more sophisticated Shapiro filter or del-
plus minus del-minus operator have been found to be no better, and often much
worse in their affects on vertical velocity. Note that a Laplacian, using no-flux
boundary conditions, conserves volume. Therefore, adding a Laplacian to the free
surface equation is trivial to implement relative to the del plus / del minus oper-
ator, which requires the specification of time dependent image points in order to
preserve volume in arbitrary geometry.
One way to apply the filter to the surface height is at the end of the barotropic time
steps. In addition to the Laplacian added to the surface height, the needs of to-
tal tracer conservation necessitate a compensating additional horizontal diffusion
be added to tracers in the surface model grid cells. Similar issues of tracer and
volume/mass compatibility were described for the cross-land mixing scheme in
Chapter 13. This approach has lost favor among the developers due to the nontriv-
ial effects it has on tracer concentration in the k = 1 level.
Another way to add a Laplacian filter is to filter the surface height within the
barotropic time loop. In this way, the surface height is continually being smoothed.
This method increases the cost of the barotropic system due to the added compu-
tations and extra communication (for parallel computers). Yet it removes the need
to concern ourselves with tracers, since compatibility is not a problem so long as
the surface height is updated consistenty with the tracer (see chapter 7). This is the
approach implemented in MOM4.
25.3. POLAR FILTERING 269
Adcroft, A., C. Hill, and J. Marshall, 1999: A new treatment of the coriolis terms
in c-grid models at both high and low resolutions. Monthly Weather Review, 127,
1928–1936.
Aris, R., 1962: Vectors, Tensors and the Basic Equations of Fluid Mechanics. Dover Pub-
lishing, 262 pp.
Beckers, J. M., 1999: On some stability properties of the discretisation of the
damped propagation of shallow-water intertia-gravity waves on the arakawa b-
grid. Ocean Modelling, 1, 53–69.
Beckmann, A., and R. Döscher, 1997: A method for improved representation of
dense water spreading over topography in geopotential–coordinate models. Jour-
nal of Physical Oceanography, 27, 581–591.
Bryan, F., 1987: Parameter sensitivity of primitive equation ocean general circula-
tion models. Journal of Physical Oceanography, 17, 970–985.
Bryan, K., 1969: A numerical method for the study of the circulation of the world
ocean. Journal Computational Physics, 4, 347–376.
Bryan, K., 1984: Accelerating the convergence to equilibrium of ocean-climate mod-
els. Journal of Physical Oceanography, 14, 666–673.
Bryan, K., 1989: The design of numerical models of the ocean circulation. J. W.
David L.T. Anderson, Ed., Oceanic circulation models: combining data and dynamics,
Vol. 44, Kluwer Academic Publishers, 465–511.
Bryan, K., and M. D. Cox, 1972: An approximate equation of state for numerical
models of the ocean circulation. Journal of Physical Oceanography, 4, 510–514.
Bryan, K., and L. J. Lewis, 1979: A water mass model of the world ocean. Journal of
Geophysical Research, 84, 2503–2517.
Bryan, K., S. Manabe, and R. C. Pacanowski, 1975: A global ocean-atmosphere cli-
mate model. part ii. the oceanic circulation. Journal of Physical Oceanography, 5,
30–46.
Campin, J.-M., and H. Goosse, 1999: Parameterization of density-driven downslop-
ing flow for a coarse-resolution ocean model in z-coordinate. Tellus, 51A, 412–
430.
272 BIBLIOGRAPHY
Greatbatch, R. J., Y. Lu, and Y. Cai, 2001: Relaxing the Boussinesq approximation in
ocean circulation models. Journal of Atmospheric and Oceanic Technology, 18, 1911–
1923.
Hallberg, R. W., 1997: Stable split time stepping schemes for large-scale ocean mod-
eling. Journal Computational Physics, 135, 54–65.
Haltiner, G. T., and R. T. Williams, 1980: Numerical Prediction and Dynamic Meteorol-
ogy. John Wiley and Sons.
Held, I. M., and V. D. Larichev, 1996: A scaling theory for horizontally homoge-
neous baroclinically unstable flow on a beta plane. Journal of Atmospheric Sciences,
53, 946–952.
Hendershott, M., 1972: The effects of solid earth deformation on global ocean tide.
Geophysical Journal of the Royal Astronomical Society, 29, 389–402.
Holloway, G., 1992: Representing topographic stress for large-scale ocean models.
Journal of Physical Oceanography, 22, 1033–1046.
Huang, R. X., 1993: Real freshwater flux as a natural boundary condition for the
salinity balance and thermohaline circulation forced by evaporation and precipi-
tation. Journal of Physical Oceanography, 23, 2428–2446.
274 BIBLIOGRAPHY
Killworth, P. D., and N. Edwards, 1999: A turbulent bottom boundary layer code for
use in numerical ocean models. Journal of Physical Oceanography, 29, 1221–1238.
Lamport, L., 1994: LATEX: A Documentation Preparation System User’s Guide and Refer-
ence Manual. Addison-Wesley.
Levitus, S., 1982: Climatological atlas of the world ocean. U.S. Government Printing
Office 13, NOAA, Washington, D.C. 163 pp.
K EY: Levitus82
A NNOTATION : 17 mircofiches
Losch, M., A. Adcroft, and J.-M. Campin, 2003: How sensitive are coarse general
circulation models to fundamental approximations in the equations of motion?
Journal of Physical Oceanography.
Lu, Y., 2001: Including non-Boussinesq effects in Boussinesq ocean circulation mod-
els. Journal of Physical Oceanography, 31, 1616–1622.
BIBLIOGRAPHY 275
Marchuk, and Kagan, 1989: Dynamics of Ocean Tides. Kluwer Academic, 327 pp.
Marotzke, J., 1991: Influence of convective adjustment on the stability of the ther-
mohaline circulation. Journal of Physical Oceanography, 21, 903–907.
Marshall, J., A. Adcroft, J.-M. Campin, and C. Hill, 2003: Atmosphere-ocean mod-
eling exploiting fluid isomorphisms. Journal of Physical Oceanography.
McDougall, T. J., D. R. Jackett, D. G. Wright, and R. Feistel, 2003: Accurate and com-
putationally efficient algorithms for potential temperature and density of seawa-
ter. Journal of Atmospheric and Oceanic Technology, 20, 730–741.
Morel, A., 1988: Optical modeling of the upper ocean in relation to its biogenous
matter content (case-i waters). JGR, 93, 10 749–10 768.
Morel, A., and D. Antoine, 1994: Heating rate within the upper ocean in relation to
its biooptical state. Journal of Physical Oceanography, 24, 1652–1665.
Morel, A., and J.-F. Berthon, 1989: Surface pigments, algal biomass profiles, and
potential production of the euphotic layer: relationship reinvestigated in view of
remote-sensing applications. Limnology and Oceanography, 34, 1542–1562.
Murray, R. J., 1996: Explicit generation of orthogonal grids for ocean models. Journal
Computational Physics, 126, 251–273.
Ohlmann, J., and D. Siegel, 2000: Ocean radiant heating. part i: Optical influences.
JPO, 30, 1833–1848.
Pacanowski, R. C., 1995: MOM2 Documentation, User’s Guide, and Reference Manual.
NOAA/Geophysical Fluid Dynamics Laboratory.
Pacanowski, R. C., K. Dixon, and A. Rosati, 1991: The GFDL Modular Ocean Model
User Guide. NOAA/Geophysical Fluid Dynamics Laboratory.
Pickard, G. L., and W. J. Emery, 1990: Descriptive Physical Oceanography. 5th ed.,
Pergamon Press, 320 pp.
Rahmstorf, S., 1993: A fast and complete convection scheme for ocean models.
Ocean Modelling, 101, 9–11.
Ray, R. D., 1998: Ocean self-attraction and loading in numerical tidal models. Ma-
rine Geodesy, 21, 181–192.
Rosati, A., and K. Miyakoda, 1988: A general circulation model for upper ocean
simulation. Journal of Physical Oceanography, 18, 1601–1626.
Semtner, A. J., 1974: An oceanic general circulation model with bottom topogra-
phy. Numerical Simulation of Weather and Climate, Technical Report No. 9, UCLA
Department of Meteorology.
Semtner, A. J., and Y. Mintz, 1977: Numerical simulation of the gulf stream and
mid-ocean eddies. Journal of Physical Oceanography, 7, 208–230.
Smith, R. D., 1999: The primitive equations in the stochastic theory of adiabatic
stratified turbulence. Journal of Physical Oceanography, 29, 1865–1880.
Smith, R. D., J. K. Dukowicz, and R. C. Malone, 1992: Parallel ocean general circu-
lation modeling. Physica D, 60, 38–61.
Smith, R. D., and J. C. McWilliams, 2002: Anisotropic horizonal viscosity for ocean
models. Ocean Modelling, 5, 129–156.
Thiele, G., and J. L. Sarmiento, 1990: Tracer dating and ocean ventilation. Journal of
Geophysical Research, 95, 9377–9391.
Veronis, G., 1973: Large scale ocean circulation. Advances in Applied Mechanics, 13,
2–92.
Visbeck, M., J. C. Marshall, T. Haine, and M. Spall, 1997: Specification of eddy trans-
fer coefficients in coarse resolution ocean circulation models. Journal of Physical
Oceanography, 27, 381–402.
BIBLIOGRAPHY 277
Wahr, J., 1998: Body tides on an elliptical, rotating, elastic and oceanless earth. Geo-
physical Journal of the Royal Astronomical Society, 64, 677–703.
Winters, K. B., and E. A. D’Asaro, 1995: Diascalar flux and the rate of fluid mixing.
Journal of Fluid Mechanics, 317, 179–193.
Wright, D. G., 1997: An equation of state for use in ocean models: Eckart’s formula
revisited. Journal of Atmospheric and Oceanic Technology, 14, 735–740.
INDEX
Adcroft et al. (1999), 108, 109, 271 Griffies et al. (2000a), 53, 108, 109, 169,
Aris (1962), 42, 271 273
Beckers (1999), 265, 271 Griffies et al. (2000b), 95, 200, 204, 211,
Beckmann and Döscher (1997), 22, 169, 273
170, 173, 174, 179, 180, 271 Griffies et al. (2001), 18, 21, 24, 75, 106,
Bryan and Cox (1972), 229, 271 109, 273
Bryan and Lewis (1979), 21, 271 Griffies (1998), 21, 273
Bryan et al. (1975), 269, 271 Griffies (2004), 14, 15, 18, 21, 22, 39, 41,
Bryan (1969), 17, 53, 74, 84, 119, 190, 44, 48, 53, 73, 74, 84–88, 102,
191, 260, 271 104–106, 108, 109, 121, 139, 149,
Bryan (1984), 106, 271 165, 176, 185, 186, 189, 195, 253,
Bryan (1987), 193, 271 273
Bryan (1989), 260, 271 Hallberg (1997), 109, 273
Campin and Goosse (1999), 22, 173– Haltiner and Williams (1980), 103, 273
180, 271 Held and Larichev (1996), 21, 273
Cox (1984), 14, 95, 120, 271 Hendershott (1972), 254, 273
Danabasoglu et al. (1996), 106, 272 Hinze (1975), 47, 273
Denman (1973), 117, 272 Holland et al. (1998), 22, 273
Dewar et al. (1998), 21, 272 Holloway (1992), 22, 257, 258, 273
Dukowicz and Smith (1994), 17, 267, Huang (1993), 88, 273
272 Jackett and McDougall (1995), 230, 273
Dukowicz et al. (1993), 17, 272 Janjić (1974), 265, 267, 274
Durran (1999), 260, 261, 272 Jerlov (1968), 116, 274
England (1995), 222, 272 Killworth and Edwards (1999), 174, 179,
Favre (1965), 47, 272 274
Feistel and Hagen (1995), 230, 231, 272 Killworth et al. (1984), 106, 274
Feistel (1993), 230, 231, 272 Killworth et al. (1991), 17, 265–268, 274
Gent and McWilliams (1990), 21, 272 Killworth (1989), 125, 274
Gent et al. (1995), 21, 190, 272 Klinger et al. (1996), 119, 120, 274
Gill (1982), 42, 46, 229, 230, 253, 272 Lamport (1994), 2, 274
Goosens et al. (1994), 2, 272 Landau and Lifshitz (1987), 42, 274
Gordon et al. (2000), 193, 272 Large et al. (1994), 21, 120, 274
Greatbatch et al. (2001), 21, 44, 48, 51, Large et al. (2001), 22, 274
104, 121, 272 Levitus (1982), 30, 274
Griffies and Hallberg (2000), 21, 266, Lin and Rood (1996), 127, 131, 274
273 Losch et al. (2003), 24, 274
Griffies et al. (1998), 21, 273 Lu (2001), 47, 274
INDEX 279
Marchuk and Kagan (1989), 253, 274 Döscher and Beckmann (1999), 22, 169,
Marotzke (1991), 125, 275 170, 179, 272
Marshall et al. (1997), 43, 119, 275
Marshall et al. (2003), 24, 275 acor parameter, 263
McDougall et al. (2002), 47, 48, 51, 275 advection velocity components
McDougall et al. (2003), 21, 230–232, discrete, 73–82
275 T-cell faces, 90–92
Messinger (1973), 265, 275 age tracers, 221–223
Morel and Antoine (1994), 115, 117, 118, available potential energy, 200
275
B-grid computational modes, 167, 265–
Morel and Berthon (1989), 118, 275
269
Morel (1988), 117, 275
B-grid variable placement, 54
Murray (1996), 20, 33, 57, 61, 64, 65,
bipolar Arctic
275
grid defined, 57
O’Brien (1986), 260, 275
fold, 57
Ohlmann and Siegel (2000), 117, 275
redundancy, 57
Pacanowski and Gnanadesikan (1998),
bit-wise exact, 34–35
20, 90, 275
boundary conditions, 30–31
Pacanowski and Griffies (1999), 14, 15,
22, 54, 56, 74, 106, 125, 130, changing processors, 34–35
189, 265, 275 checkerboard mixing, 269
Pacanowski and Philander (1981), 21, checkerboard mode, 265
120, 276 chlorophyll, 116–118
Pacanowski et al. (1991), 14, 275 communicating with developers, 17
Pacanowski (1995), 14, 275 commutator, 43
Paulson and Simpson (1977), 116, 117, compatibility between mass and tracer,
276 158, 165
Pedlosky (1987), 42, 276 constant volume ocean, 142–144
Pickard and Emery (1990), 203, 276 convective adjustment, 125
Press et al. (1992), 125, 276 coordinates, 42–43
Rahmstorf (1993), 120, 125, 276 cartesian and spherical, 42
Ray (1998), 254, 276 general orthogonal
Rosati and Miyakoda (1988), 116, 276 infinitesimal distances, 42–43
Semtner and Mintz (1977), 266, 276 key formulae, 42–43
Semtner (1974), 87, 276 metric tensor, 42–43
Smith and McWilliams (2002), 22, 276 partial derivatives, 42–43
Smith et al. (1992), 17, 276 Coriolis force, 259–263
Smith (1999), 47, 276 coupling tracer to velocity cells, 73–82
Thiele and Sarmiento (1990), 222, 276 cross-land mixing, 157–167
Veronis (1973), 42, 276 null modes, 167, 269
Visbeck et al. (1997), 21, 276 partial cells and free surface, 163–
Wahr (1998), 253, 276 165
Webb et al. (1998), 54, 277 tracer and mass compatibility, 158
Winters and D’Asaro (1995), 199, 201, tracer and volume compatibility,
204, 277 158
Winters et al. (1995), 199, 201, 277
Winton et al. (1998), 169, 173, 174, 277 data override, 30–31
Wright (1997), 231, 277 density
280 INDEX