1
Stiftung Alfred-Wegener-Institut ur Polar- und Meeresforschung Adjoint Sensitivity of an Ocean General Circulation Model to Bottom Topography Martin Losch 1 and Patrick Heimbach 2 1 Alfred-Wegener-Institut f ¨ ur Polar- und Meeresforschung, Bremerhaven, Germany, email: [email protected] 2 Dep. Earth, Atmospheric and Planetary Sciences, Massachusetts Institute of Technology, Cambridge, USA, email: [email protected] 1 Overview Numerical solutions of ocean general circulation models are determined by many different pa- rameters. So far, state estimation based on inverse methods, such as that of the ECCO consor- tium, use surface boundary conditions and, for integrations shorter than the time for baroclinic adjustement processes to complete, initial conditions as canonical control variables. Other pa- rameters, for example diffusivities, lateral boundary conditions (free-slip, no-slip, etc.), and bottom topography, are formally assumed to be known. However, it is not clear what their “correct” values in coarse resolution 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 ambiguous and may add numerical artifacts to the ocean model’s solution. Here, we extend the approach of Losch and Wunsch (2003), who use topography as a control variable in a simpler model, to a full general circulation model Losch and Heimbach (2006). Figure 1: a) 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. b) Topography of numerical model in this study. -500 0 500 0 o 90 o E 180 o W 90 o W 0 o 60 o S 30 o S 0 o 30 o N 60 o N a) difference between ETOPO2 and ETOPO5 [m] -6000 -5000 -4000 -3000 -2000 -1000 0 b) model topography [m] 0 o 90 o E 180 o W 90 o W 0 o 60 o S 30 o S 0 o 30 o N 60 o N 2 MITgcm and its adjoint The M.I.T. General Circulation Model (MIT- gcm) is rooted in a general purpose grid-point algorithm that solves the Boussinesq form of the Navier-Stokes equations for an incompress- ible fluid, hydrostatic or fully non-hydrostatic, in a curvilinear framework (in the present con- text on a three-dimensional longitude (λ), lat- itude (ϕ), depth (H ) grid. The algorithm is described in Marshall et al. (1997) (for online documentation and access to the model code, see MITgcm Group, 2002)). The MITgcm has been adapted for use with the Tangent linear and Adjoint Model Com- piler (TAMC), and its successor TAF (Trans- formation of Algorithms in Fortran, Gier- ing and Kaminski, 1998). Efficient (w.r.t. CPU/memory), exact (w.r.t. the model’s tran- sient 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, 2005). 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 complexities to the problem: 1. expressions involving the product of an el- ement of the model state and topographic masks are now quadratic w.r.t. algorithmic differentiation; 2. the elliptic operator to solve for the sur- face pressure/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 a) ∂φ/∂τ x [Sv m 2 /N] 0 o 90 o E 180 o W 90 o W 0 o 60 o S 30 o S 0 o 30 o N 60 o N -2 -1 0 1 2 b) ∂φ/∂τ y [Sv m 2 /N] 0 o 90 o E 180 o W 90 o W 0 o 60 o S 30 o S 0 o 30 o N 60 o N -2 -1 0 1 2 c) ∂φ/Q [10 7 Sv s 3 /m 2 ] 0 o 90 o E 180 o W 90 o W 0 o 60 o S 30 o S 0 o 30 o N 60 o N -1 -0.5 0 0.5 1 d) ∂φ/H [10 -3 Sv/m] 0 o 90 o E 180 o W 90 o W 0 o 60 o S 30 o S 0 o 30 o N 60 o N 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/∂τ x , ∂φ/∂τ x ), surface buoyancy fluxes (∂φ/∂Q, ∂φ/∂EmP ), and bottom topography (∂φ/∂H ). The volume transport is “measured” at the end of a 100 year spin-up integration starting from rest. 4 Atlantic Overturning Streamfunction at 24 N -2 -1 0 1 2 a) ∂φ/∂τ x [Sv m 2 /N] 0 o 90 o E 180 o W 90 o W 0 o 60 o S 30 o S 0 o 30 o N 60 o N -2 -1 0 1 2 b) ∂φ/∂τ y [Sv m 2 /N] 0 o 90 o E 180 o W 90 o W 0 o 60 o S 30 o S 0 o 30 o N 60 o N -2 -1 0 1 2 c) ∂φ/Q [10 7 Sv s 3 /m 2 ] 0 o 90 o E 180 o W 90 o W 0 o 60 o S 30 o S 0 o 30 o N 60 o N -1 -0.5 0 0.5 1 d) ∂φ/H [10 -3 Sv/m] 0 o 90 o E 180 o W 90 o W 0 o 60 o S 30 o S 0 o 30 o N 60 o N Figure 3: Adjoint sensitivity of meridional overturning strength φ at 24 N to surface stresses, surface buoyancy fluxes and bottom topography. The overturning strength is “measured” at the end of a 100 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 scale and spatial resolution 1 year 0 o 90 o E 180 o W 90 o W 0 o 60 o S 30 o S 0 o 30 o N 60 o N 10 years 0 o 90 o E 180 o W 90 o W 0 o 60 o S 30 o S 0 o 30 o N 60 o N 500 years 0 o 90 o E 180 o W 90 o W 0 o 60 o S 30 o S 0 o 30 o N 60 o N ∂φ/H [10 -3 Sv/m] -1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1 1000 years 0 o 90 o E 180 o W 90 o W 0 o 60 o S 30 o S 0 o 30 o N 60 o N 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 sensitivity 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, 500, and 1000 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 o 60 o E 120 o E 180 o W 120 o W 60 o W 0 o 60 o S 30 o S 0 o 30 o N 60 o N 1 year 0 o 60 o E 120 o E 180 o W 120 o W 60 o W 0 o 60 o S 30 o S 0 o 30 o N 60 o N 10 years 0 o 60 o E 120 o E 180 o W 120 o W 60 o W 0 o 60 o S 30 o S 0 o 30 o N 60 o N ∂φ/H [10 -4 Sv/m] -5 -4 -3 -2 -1 0 1 2 3 4 5 10 years 0 o 60 o E 120 o E 180 o W 120 o W 60 o W 0 o 60 o S 30 o S 0 o 30 o N 60 o N Figure 5: The adjoint sensitivity of volume transport through the Drake Passage to bottom topography is indeed sensitive to specifics of the model configuration 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 experiment. 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 τ x H u 0.1 N/m 2 100 m ∂φ/∂u 2 Sv m 2 /N -2 · 10 -3 Sv/m φ 0.2 Sv -0.2 Sv location Drake Passage 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 topogra- phy 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 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. (2005). An efficient exact adjoint of the parallel MIT general circulation model, generated via automatic differentiation. Future Genera- tion Computer Systems (FGCS), 21(8):1356– 1371. Losch, M. and Heimbach, P. (2006). Adjoint sen- sitivity of an ocean general circulation model to bottom topography. J. Phys. Oceanogr., in press. Losch, M. and Wunsch, C. (2003). Bottom to- pography as a control parameter in an ocean circulation model. J. Atmos. Ocean. Technol., 20(11):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):5733–5752. MITgcm Group (2002). MITgcm Re- lease 1 Manual. (online documentation), MIT/EAPS, Cambridge, MA 02139, USA. http://mitgcm.org/sealion/ online_documents/manual.html.

Adjoint Sensitivity of an Ocean General Circulation Model to Bottom Topography … · 2011. 9. 28. · Adjoint Sensitivity of an Ocean General Circulation Model to Bottom Topography

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Adjoint Sensitivity of an Ocean General Circulation Model to Bottom Topography … · 2011. 9. 28. · Adjoint Sensitivity of an Ocean General Circulation Model to Bottom Topography

Stiftung Alfred-Wegener-Institutfur Polar- und Meeresforschung

Adjoint Sensitivity of an Ocean General CirculationModel to Bottom Topography

Martin Losch 1 and Patrick Heimbach 2

1Alfred-Wegener-Institut fur Polar- und Meeresforschung , Bremerhaven, Germany, email: [email protected] e2Dep. Earth, Atmospheric and Planetary Sciences, Massachus etts Institute of Technology, Cambridge, USA, email: heimb [email protected]

1 Overview

Numerical solutions of ocean general circulation models are determined by many different pa-rameters. So far, state estimation based on inverse methods, such as that of the ECCO consor-tium, use surface boundary conditions and, for integrations shorter than the time for baroclinicadjustement processes to complete, initial conditions as canonical control variables. Other pa-rameters, for example diffusivities, lateral boundary conditions (free-slip, no-slip, etc.), andbottom topography, are formally assumed to be known. However, it is not clear what their“correct” values in coarse resolution models are. Bottom topography, for example is not knownaccurately in large regions of the ocean. Even where it is known, its representation on a coarsegrid is ambiguous and may add numerical artifacts to the ocean model’s solution. Here, weextend the approach of Losch and Wunsch (2003), who use topography as a control variable ina simpler model, to a full general circulation model Losch and Heimbach (2006).

Figure 1: a) Difference between two sea floor topographies (ETOPO2 minus ETOPO5),smoothed with a 1◦ radial block average. Not only are the differences locally larger than1000 m, but also is ETOPO2 systematically shallower in partsof the Southern Ocean, e.g.,south of the Australian continent. b) Topography of numerical model in this study.

−500

0

500

0o 90oE 180oW 90oW 0o

60oS

30oS

0o

30oN

60oN

a) difference between ETOPO2 and ETOPO5 [m]

−6000

−5000

−4000

−3000

−2000

−1000

0b) model topography [m]

0o 90oE 180oW 90oW 0o

60oS

30oS

0o

30oN

60oN

2 MITgcm and its adjointThe M.I.T. General Circulation Model (MIT-gcm) is rooted in a general purpose grid-pointalgorithm that solves the Boussinesq form ofthe Navier-Stokes equations for an incompress-ible fluid, hydrostatic or fully non-hydrostatic,in a curvilinear framework (in the present con-text on a three-dimensional longitude (λ), lat-itude (ϕ), depth (H) grid. The algorithm isdescribed in Marshall et al. (1997) (for onlinedocumentation and access to the model code,see MITgcm Group, 2002)).The MITgcm has been adapted for use withthe Tangent linear and Adjoint Model Com-piler (TAMC), and its successor TAF (Trans-formation of Algorithms in Fortran, Gier-ing and Kaminski, 1998). Efficient (w.r.t.CPU/memory), exact (w.r.t. the model’s tran-

sient state) derivative code can be generatedfor up-to-date versions of the MITgcm and itsnewly developed packages in a wide range ofconfigurations (Heimbach et al., 2002, 2005).In addition to the general hurdles that need tobe tackled for efficient exact adjoint code gen-eration, inclusion of bottom topography as acontrol variable added further complexities tothe problem:

1. expressions involving the product of an el-ement of the model state and topographicmasks are now quadratic w.r.t. algorithmicdifferentiation;

2. the elliptic operator to solve for the sur-face pressure/height field now looses its self-adjoint property, and needs to be explicitlyadjointed as well.

3 Drake Passage Transport

−2 −1 0 1 2

a) ∂φ/∂τx [Sv m2/N]

0o 90oE 180oW 90oW 0o 60oS

30oS

0o

30oN

60oN

−2 −1 0 1 2

b) ∂φ/∂τy [Sv m2/N]

0o 90oE 180oW 90oW 0o 60oS

30oS

0o

30oN

60oN

−2 −1 0 1 2

c) ∂φ/∂Q [107Sv s3/m2]

0o 90oE 180oW 90oW 0o 60oS

30oS

0o

30oN

60oN

−1 −0.5 0 0.5 1

d) ∂φ/∂H [10−3Sv/m]

0o 90oE 180oW 90oW 0o 60oS

30oS

0o

30oN

60oN

Figure 2: The Drake Passage is a sensitive “choke point” in the Southern Ocean. We show the adjoint sensitivity ofvolume transport through the Drake Passageφ 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 100 year spin-upintegration starting from rest.

4 Atlantic Overturning Streamfunction at 24 ◦ N

−2 −1 0 1 2

a) ∂φ/∂τx [Sv m2/N]

0o 90oE 180oW 90oW 0o 60oS

30oS

0o

30oN

60oN

−2 −1 0 1 2

b) ∂φ/∂τy [Sv m2/N]

0o 90oE 180oW 90oW 0o 60oS

30oS

0o

30oN

60oN

−2 −1 0 1 2

c) ∂φ/∂Q [107Sv s3/m2]

0o 90oE 180oW 90oW 0o 60oS

30oS

0o

30oN

60oN

−1 −0.5 0 0.5 1

d) ∂φ/∂H [10−3Sv/m]

0o 90oE 180oW 90oW 0o 60oS

30oS

0o

30oN

60oN

Figure 3: Adjoint sensitivity of meridional overturning strengthφ at 24◦N to surface stresses, surface buoyancy fluxesand bottom topography. The overturning strength is “measured” at the end of a 100 year spin-up integration starting fromrest. Other than for the Drake Passage transport, the largest sensitivities are found close to the location of the objectivefunction.

5 Sensitivity to temporal scale and spatial resolution1 year

0o 90oE 180oW 90oW 0o 60oS

30oS

0o

30oN

60oN

10 years

0o 90oE 180oW 90oW 0o 60oS

30oS

0o

30oN

60oN

500 years

0o 90oE 180oW 90oW 0o 60oS

30oS

0o

30oN

60oN ∂φ/∂

H [1

0−3 S

v/m

]

−1

−0.8

−0.6

−0.4

−0.2

0

0.2

0.4

0.6

0.8

1

1000 years

0o 90oE 180oW 90oW 0o 60oS

30oS

0o

30oN

60oN

Figure 4: The adjoint sensitivity to bottom topography in Fig.2 depends only marginallyon the length of the spin-up integration. The main patterns of sensitivity near the DrakePassage, the Kerguelen Plateau, the Macquarie Ridge and theIndonesian Throughflow arerobust with respect to length of the spin-up integration; shown are runs of 1, 10, 500, and1000 years. It is worth noting that with increasing integration length the Drake Passagetransport becomes sensitive to the topography of the North Atlantic (see also Fig.2).

1 year

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

60oS

30oS

0o

30oN

60oN

1 year

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

60oS

30oS

0o

30oN

60oN

10 years

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

60oS

30oS

0o

30oN

60oN ∂φ/∂

H [1

0−4 S

v/m

]

−5

−4

−3

−2

−1

0

1

2

3

4

5

10 years

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

60oS

30oS

0o

30oN

60oN

Figure 5: The adjoint sensitivity of volume transport through the Drake Passage to bottomtopography is indeed sensitive to specifics of the model configuration such as horizontaland vertical resolution. For a configuration with twice the resolution (2 by 2 degrees)than in the previous experiments, the spatial pattern of theadjoint sensitivity is far moredetailed than in the coarse resolution experiment (Fig.4),but the large scale patterns remainsimilar. The Antarctic Circumpolar Current is steered around the Kerguelen Plateau in thisexperiment, thus reducing the sensitivity seen in the coarse resolution experiment. Thevolume transport is “measured” at the end of a 10 year spin-upintegration starting fromrest.

6 Discussion

One can assess the relative importance of different controlvariables by multiplying the respective gradient by their apriori uncertainty estimates, i.e.

∆φ =∂φ

∂u∆u

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

control variableu τx H

∆u 0.1N/m2 100 m

∂φ/∂u 2 Sv m2/N −2 · 10

−3Sv/m

∆φ 0.2 Sv −0.2 Svlocation Drake PassageKerguelen Plateau

This information can be used to systematically adjust thegeometry of a coarse (w.r.t. to the actual topography)general circulation model. Eventually, bottom topogra-phy and further “unorthodox” parameters such a lateralboundary conditions (no slip, free slip) will be includedin the control vector of a state estimation problem.

ReferencesGiering, R. and Kaminski, T. (1998). Recipes for

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

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

Automatic generation of efficient adjoint codefor 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 ofLecture Notes in Com-puter Science, pages 1019–1028. Springer-

Verlag, Berlin (Germany).

Heimbach, P., Hill, C., and Giering, R. (2005).An efficient exact adjoint of the parallelMIT general circulation model, generated viaautomatic differentiation. Future Genera-tion Computer Systems (FGCS), 21(8):1356–

1371.

Losch, M. and Heimbach, P. (2006). Adjoint sen-sitivity of an ocean general circulation modelto bottom topography.J. Phys. Oceanogr., inpress.

Losch, M. and Wunsch, C. (2003). Bottom to-

pography as a control parameter in an oceancirculation model.J. Atmos. Ocean. Technol.,20(11):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):5733–5752.

MITgcm Group (2002). MITgcm Re-lease 1 Manual. (online documentation),MIT/EAPS, Cambridge, MA 02139, USA.http://mitgcm.org/sealion/online_documents/manual.html.