• Keine Ergebnisse gefunden

Towards including bottom topography as a control variable in global ocean state estimation

N/A
N/A
Protected

Academic year: 2022

Aktie "Towards including bottom topography as a control variable in global ocean state estimation"

Copied!
1
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Stiftung Alfred-Wegener-Institut f ¨ur Polar- und Meeresforschung

Towards including bottom topography as a control variable in global ocean state estimation

Martin Losch

1

and Patrick Heimbach

2

1Alfred-Wegener-Institut f ¨ur Polar- und Meeresforschung, Bremerhaven, Germany, email: mlosch@awi-bremerhaven.de

2Dep. Earth, Atmospheric and Planetary Sciences, Massachusetts Institute of Technology, Cambridge, USA, email: heimbach@mit.edu

1 Overview

Numerical solutions of ocean general circulation models are determined by many different parameters. So far, state esti- mation based on inverse methods, such as that of the ECCO consortium, use surface boundary conditions and, for in- tegrations shorter than the time for baroclinic adjustement processes to complete, initial conditions as canonical con- trol variables. Other parameters, for example diffusivities, lateral boundary conditions (free-slip, no-slip, etc.), and bot- tom topography, are formally assumed to be known. How- ever, it is not clear what their “correct” values in coarse res- olution models are. Bottom topography, for example is not known accurately in large regions of the ocean. Even where it is known, its representation on a coarse grid is ambigu- ous and may add numerical artifacts to the ocean model’s solution. Here, we extend the approach of Losch and Wun-

sch (2003), who use topography as a control variable in a simpler model, to a full general circulation model.

−1000

−500 0 500 1000

0o 60oE 120oE 180oW 120oW 60oW 0o 60oS

30oS 0o 30oN 60oN

Figure 1: Difference between two sea floor topographies (ETOPO2 minus ETOPO5), smoothed with a 1 radial block average. Not only are the differences locally larger than 1000 m, but also is ETOPO2 systematically shallower in parts of the Southern Ocean, e.g., south of the Australian continent.

2 MITgcm and its adjoint

The M.I.T. General Circulation Model (MITgcm) is rooted in a general purpose grid-point algorithm that solves the Boussinesq form of the Navier-Stokes equations for an in- compressible fluid, hydrostatic or fully non-hydrostatic, in a curvilinear framework (in the present context on a three- dimensional longitude (λ), latitude (ϕ), depth (H) grid. The algorithm is described in Marshall et al. (1997) (for online documentation and access to the model code, see Adcroft et al., 2002)).

The MITgcm has been adapted for use with the Tangent lin- ear and Adjoint Model Compiler (TAMC), and its succes- sor TAF (Transformation of Algorithms in Fortran, Giering and Kaminski, 1998). Efficient (w.r.t. CPU/memory), ex- act (w.r.t. the model’s transient state) derivative code can

be generated for up-to-date versions of the MITgcm and its newly developed packages in a wide range of configurations (Heimbach et al., 2002, 2004).

In addition to the general hurdles that need to be tackled for efficient exact adjoint code generation, inclusion of bottom topography as a control variable added further complexities to the problem:

1. expressions involving the product of an element of the model state and topographic masks are now quadratic w.r.t. algorithmic differentiation;

2. the elliptic operator to solve for the surface pres- sure/height field now looses its self-adjoint property, and needs to be explicitly adjointed as well.

3 Drake Passage Transport

−2 −1 0 1 2

∂φ/∂τx [Sv m2/N]

0 100 200 300

−50 0 50

−2 −1 0 1 2

∂φ/∂τy [Sv m2/N]

0 100 200 300

−50 0 50

−15 −10 −5 0 5 10 15

∂φ/Q [10−3Sv m2/W]

0 100 200 300

−50 0 50

−10 −5 0 5 10

∂φ/EmP [106Sv s/m]

0 100 200 300

−50 0 50

−2 −1 0 1 2

∂φ/H [10−3Sv/m]

0 100 200 300

−50 0 50

−6 −5 −4 −3 −2 −1 0

global4x4/run1000y02

−H(x,y) [km]

0 100 200 300

−50 0 50

Figure 2: The Drake Passage is a sensitive “choke point” in the Southern Ocean. We show the adjoint sensitivity of volume transport through the Drake Passage φ to surface stresses (∂φ/∂τx, ∂φ/∂τx), surface buoyancy fluxes (∂φ/∂Q,

∂φ/∂EmP), and bottom topography (∂φ/∂H). The volume transport is “measured” at the end of a 1000 year spin-up integration starting from rest.

4 Atlantic Overturning Streamfunction at 24

N

−5 0 5

∂φ/∂τx [Sv m2/N]

0 100 200 300

−50 0 50

−5 0 5

∂φ/∂τy [Sv m2/N]

0 100 200 300

−50 0 50

−30 −20 −10 0 10 20 30

∂φ/Q [10−3Sv m2/W]

0 100 200 300

−50 0 50

−20 −10 0 10 20

∂φ/EmP [106Sv s/m]

0 100 200 300

−50 0 50

−2 −1 0 1 2

∂φ/H [10−3Sv/m]

0 100 200 300

−50 0 50

−6 −5 −4 −3 −2 −1 0

global4x4/run1000y06

−H(x,y) [km]

0 100 200 300

−50 0 50

Figure 3: Adjoint sensitivity of meridional overturning strength φ at 24N to surface stresses, surface buoyancy fluxes and bottom topography. The overturning strength is “measured” at the end of a 1000 year spin-up integration starting from rest. Other than for the Drake Passage transport, the largest sensitivities are found close to the location of the objective function.

5 Sensitivity to temporal and spatial scale and resolution

1 year

0 50 100 150 200 250 300 350

−50 0 50

10 years

0 50 100 150 200 250 300 350

−50 0 50

100 years

0 50 100 150 200 250 300 350

−50 0 50

−2

−1.5

−1

−0.5 0 0.5 1 1.5 2

∂φ/H [10−3 Sv m2/N]

500 years

0 50 100 150 200 250 300 350

−50 0 50

Figure 4: The adjoint sensitivity to bottom topography in Fig.2 depends only marginally on the length of the spin-up integration. The main patterns of sen- sitivity near the Drake Passage, the Kerguelen Plateau, the Macquarie Ridge and the Indonesian Throughflow are robust with respect to length of the spin- up integration; shown are runs of 1, 10, 100, 500 years. It is worth noting that with increasing integration length the Drake Passage transport becomes sensitive to the topography of the North Atlantic (see also Fig.2).

1 year

0 50 100 150 200 250 300 350

−50 0 50

−2

−1 0 1 2 10 years

0 50 100 150 200 250 300 350

−50 0 50

Figure 5: The adjoint sensitivity of volume transport through the Drake Pas- sage to bottom topography is indeed sensitive to specifics of the model con- figuration such as horizontal and vertical resolution. For a configuration with twice the resolution (2 by 2 degrees) than in the previous experiments, the spatial pattern of the adjoint sensitivity is far more detailed than in the coarse resolution experiment (Fig.4), but the large scale patterns remain similar. The Antarctic Circumpolar Current is steered around the Kerguelen Plateau in this experiment, thus reducing the sensitivity seen in the coarse resolution exper- iment. The volume transport is “measured” at the end of a 10 year spin-up integration starting from rest.

6 Discussion

One can assess the relative importance of different control variables by multiplying the respective gradient by their a priori uncertainty estimates, i.e.

∆φ = ∂φ

∂u ∆u

with u a control variable element and ∆u some a priori uncertainty estimate. The following table illustrates that in places, the effect of bottom topography sensitivities are of similar magnitude as that of wind stress sensitivities (c.f. Fig.2).

control variable u ∆u ∂φ/∂u ∆φ location

τx 0.1 N/m2 2 Sv m2/N 0.2 Sv over Drake Passage

H 100 m −2 · 103 Sv/m −0.2 Sv over Kerguelen Plateau

This information can be used to systematically adjust the geometry of a coarse (w.r.t.

to the actual topography) general circulation model. Eventually, bottom topography and further “unorthodox” parameters such a lateral boundary conditions (no slip, free slip) will be included in the control vector of a state estimation problem.

References

Adcroft, A., Campin, J.-M., Heimbach, P., Hill, C., and Marshall, J. (2002). MIT- gcm Release 1 Manual. (online documenta-

tion) , MIT/EAPS, Cambridge, MA 02139, USA. http://mitgcm.org/sealion/

online_documents/manual.html. Giering, R. and Kaminski, T. (1998). Recipes for

adjoint code construction. ACM Transactions on Mathematical Software, 24:437–474.

Heimbach, P., Hill, C., and Giering, R. (2002).

Automatic generation of efficient adjoint code for a parallel Navier-Stokes solver. In J.J.

Dongarra, P.M.A. Sloot and C.J.K. Tan, edi- tor, Computational Science – ICCS 2002, vol- ume 2331, part 3 of Lecture Notes in Com-

puter Science, pages 1019–1028. Springer- Verlag, Berlin (Germany).

Heimbach, P., Hill, C., and Giering, R. (2004).

An efficient exact adjoint of the parallel MIT general circulation model, generated via au- tomatic differentiation. Future Generation

Computer Systems, in press.

Losch, M. and Wunsch, C. (2003). Bottom to- pography as a control variable in an ocean model. J. Atmos. Ocean. Technol., 20:1685–

1696.

Marshall, J., Hill, C., Perelman, L., and Adcroft, A. (1997). Hydrostatic, quasi-hydrostatic and nonhydrostatic ocean modeling. J. Geo- phys. Res., 102, C3:5,733–5,752.

Referenzen

ÄHNLICHE DOKUMENTE

\~here xperp is the space variable defined perpendicular to the sand wave crest, k is the wave number vector of short gravity waves, ko i the lower limit of the

(The case of coordinate in somewhat similar problems was studied in [10, 11], where the feedback assigned points were not time moments but coordinates of point heat sources on a

cedure fits a normal distribution to the three values (high, central and low) that resulted from expert discussions, with 90 percent o f the cases lying between the high

Noting that various commentators have questioned both premises, Gillman tests this by examining seven different measures of the general rate of profit (total

Blanchard and Kahn ’s (1980) solution for a linear rational expectations model with one state variable and one control variable: the correct formula.. Robert Kollmann

With the help of an adjoint model, it is shown that scalar diagnostics or objective functions in a coarse resolution model, such as the transport through Drake Passage, the strength

In addition to the general hurdles that need to be tackled for efficient exact adjoint code gen- eration, inclusion of bottom topography as a control variable added further

Difference between true depth and estimated depth (in meters, top) and estimated error reduction (in percent, relative to the initial error estimate of 200m, bottom); noisy