global stokes drift and - deepsthe stokes drift can be defined as the mean temporal and spatial...

1
Gaussian RBFs are used to define an interpolated solution (in this case, wave spectral density) as where ______________________ is the RBF shape parameter, and ‘N’ is the total number of nodes. The wave action balance equation is then projected on sphere as and a unique n x n stencil (nN) is constructed for each node. For deep water and without currents or source terms, the numerical scheme (with hyperviscosity) becomes where c g is the group velocity, !’ represents element multiplication, and D N is a sparse N x N differentiation matrix. This can be solved using a parallelized explicit time-stepping method. VII. Unstructured Node Prototype The previously mentioned disadvantages (pole singularity and shrinking grid cells) are an artifact of representing 3D spherical data on a 2D plane and a decreasing time step due to the CFL condition. One possibility being explored is the use of unstructured nodes on a 3D sphere (both spatial and wavenumber) to remove the singularity and spread the computational costs equally. Radial basis functions (RBFs) use an arbitrary node layout and are well-suited to solving convective PDEs on a sphere due to their algorithmic simplicity and spectral accuracy. However, the method is computationally expensive when scaled to a large number of 'N' nodes. To circumvent this, the RBF generated finite-difference (RBF-FD) method uses large stencils to generate differential weights to re- produce RBFs (rather than polynomials). This meth- od scales as O(N) per time step and is parallelizable. Figure 7: Example of an unstructured node layout to solve the wave action balance equation on an aquaplanet. t N + Pk · Px N + Px · Pk N = P(Source Terms) N (α,t) t=t 0 = N i=1 λ i φ(α α i ), for φ(r )= e (εr) 2 and ε,r R + , t N = c g x D x N N c g y D y N N c g z D z N N Adrean Webb Dept. of Applied Math, CIRES [email protected] Global Stokes Drift and Climate Wave Modeling Baylor Fox-Kemper Dept. of Atm. and Oceanic Sciences, CIRES [email protected] The Stokes drift velocity - the Lagrangian-average fluid velocity induced from wave action - is an important vector component that appears often in wave-averaged dynamics, such as Langmuir turbulence. However, accuracy and data coverage remain challenges in estimating Stokes drift globally and recent research (Hanley et al., 2010, Webb & Fox-Kemper, 2011) has shown that using atmospheric data and the assumption of wave equilibration is not trustworthy, as often the wave state is dominated by developing or remotely-generated swell conditions. Figure 1: Example of the global variability of Stokes drift (created using WWATCH III and CORE2 winds). II. Importance for Climate Research Stokes drift is of interest in climate research due to its dominant role in determining the strength of Langmuir turbulence - surface mixing due to the interaction of wind and waves. Preliminary work (Fox-Kemper et al, 2011) to include this unresolved turbulence in climate models, has shown the potential to correct a well-known, shallow mixed-layer bias in the Southern Ocean. Figure 2: (a) Example of Langmuir turbulence (Deep Water Horizon spill), (b) Southern Ocean mixed-layer depth results from a preliminary parametrization of Langmuir turbulence (NCAR CCSM 3.5) npr.org VI. Need for a Climate Wave Model Unfortunately, all conventional, structured latitude-longitude-grid wave-models have a disadvan- tage in that they are singular at the poles and the grid cell sizes shrink appreciably with higher latitudes. This is a problem for climate simulations since the Arctic may be ice-free in the near future and performance of the model will deteriorate as the northern boundary is moved higher. Overcoming these problems and improving performance for long model runs (without compromising climate-scale statistics), remains a challenge. V. Coupling Wave and Circulation Models For climate research purposes, it is advantageous to use a prognostic 2D wave field - not only to calculate Langmuir turbulence but also to improve atmospheric and oceanic fluxes. As such, NCAR is currently adding a modified form of WWATCH III (v3.14) to the development code of the Community Earth System Model (CESM). The wave module component will be linked to the atmosphere-ocean coupler on 30 minute intervals. Since it is not uncommon for climate runs to span several hundred years, a high model throughput of 30 or more simulated years per computational day is ideal and early benchmarking shows that this is achievable if a coarsened grid of 3.2 x 4 deg is used with a northern and southern latitude cutoff of 70 deg. Figure 6: WWATCH III grid performance with early benchmarking targets on two different machines: (a) NASA Pleaides, (b) NCAR Bluefire. Performance is measured in number of simulated years per day of running. 30 simulated years per day or greater is necessary to couple with ECSM α =(x,y,z,k x ,k y ,k z ), ε VIII. Conclusion In computations where the accuracy of Stokes drift is important, it is essential to use either the full 2D wave spectra or 1D spectra with an empirically-derived spreading function. Currently, WWATCH III is being coupled to the NCAR CESM to add Langmuir turbulence and improve air- sea fluxes. In addition, a prototype climate wave model is being developed for long model runs and ice-free conditions in the Arctic. While there are still major challenges to overcome (such as boundaries and parallelization), this unstructured node approach shows promise. III. Leading-order Stokes Drift The Stokes drift can be defined as the mean temporal and spatial difference between the Eulerian and Lagrangian velocities for a finite period and length scale. Using a series expansion and a linear wave decomposition, the leading-order Stokes drift can be rewritten in spectral density form as Notice that the leading order Stokes drift is a vector quantity whose magnitude depends both on the directional components of the wave field and the directional spread of wave energy for each component. To illustrate how different directional wave components might affect the magnitude of Stokes drift, consider a pair of monochromatic waves traveling with an incident angle !’ from the y-axis (Fig. 3). Then the magnitude of Stokes drift will depend on sin !’. Figure 3: An example of how the directional components of a wave field can affect the magnitude of the Stokes drift. u S 16π 3 g 0 π π cos θ , sin θ , 0 f 3 S f θ (f, θ ) e 8π 2 f 2 g z dθ df !IV. One-dimensional Stokes Drift Commonly, a unidirectional wave approxi- mation (|H|=1) is used to simplify the 1D Stokes drift. However, this tends to over- estimate the full leading-order Stokes drift by 33% (in a 3rd generation wave model) and is not recommended (Webb & Fox-Kemper, 2011). A better approximation can be found by using an empirically-determined frequency- dependent spread function for " f . To illustrate, a ‘directional loss’ is plotted in Fig. 5 using a spread function derived by Donelan et al. (1985). 0.95,0.93 0.56,0.78 1.6,0.78 0 0.5 1 1.5 2 0.7 0.8 0.9 1 f f p H f f p Often only 1D spectral data, or even just spectral moments, are available and this poses a challenge to accurately estimate Stokes drift (see Fig. 4 for example). 0 10 30 50 60 80 90 110 100 70 40 20 0 0.05 0.10 0.15 0.20 0.25 0.30 0.35 Frequency (Hz) Wave Spectral Density (m 2 / Hz) 20.6 m/s 18 m/s 15.4 m/s 12.9 m/s 10.3 m/s As illustrated previously, 1D estimates will tend to overestimate the full leading-order Stokes drift when there is more than one directional component. However, the error due to wave spreading can be minimized by using an empirical spreading function. Consider splitting the 2D spectral density as where " f is the ‘directional distribution’. Then a directional loss H due to wave energy being directed along other directions other than the dominant direction, can be defined as and the 1D Stokes drift approximation can be written as 0 π π S f θ (f, θ ) dθ df = 0 π π φ f (f, θ ) S f (f ) dθ df = 0 S f (f ) df, H(f )= π π cos θ , sin θ , 0 φ f (f, θ ) dθ Figure 4: Examples of wave spectra: (a) 2D spectra generated by WWATCH III, (b) 1D Pierson and Moskowitz observational spectra (Stewart, Intro. to Physical Oceanography). Figure 5: An example of directional spread loss using an empirically-determined spread function from Donelan et al. (1985). Acknowledgements Thanks to Natasha Flyer at NCAR for advice on implementing the RBF-FD method. Work supported by NASA NNX09AF38G, NSF 0934737, CIRES, UCAR, and CU-Boulder. References Fornberg, B., Lehto, E., 2011. Stabilization of RBF-generated finite difference methods for convective PDEs, Journal of Computational Physics, Volume 230, Issue 6. Donelan, M.A., Hamilton, J., Hui, W.H., 1985. Directional spectra of wind-generated waves. Philosophical Transactions of the Royal Society of London Series A - Mathematical Physical and Enineering Sciences 315, 509-562. Fox-Kemper, B., Webb, A., Baldwin-Stevens, E., Danabasoglu, G., Hamlington, B., Large, WG., Peacock, S., in preparation. Global climate model sensitivity to estimated Langmuir Mixing. Ocean Modelling. Hanley, K.E., Belcher, S.E., Sullivan, P.P., 2010. A global climatology of wind–wave interaction. Journal of Physical Oceanography 40, 1263–1282. Flyer, N., Wright, G., 2009. A Radial Basis Function Method for the Shallow Water Equations on a Sphere. Proceedings of the Royal Society A - Mathematical Physical and Engineering Sciences. Webb, A., Fox-Kemper, B., 2011. Wave spectral moments and Stokes drift estimation. Ocean Modelling. I. Introduction u s 16π 3 g 0 H(f ) f 3 S f (f ) e 8π 2 f 2 g z df

Upload: others

Post on 25-May-2020

9 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Global Stokes Drift and - DEEPSThe Stokes drift can be defined as the mean temporal and spatial difference between the Eulerian and Lagrangian velocities for a finite period and length

Gaussian RBFs are used to define an interpolated solution (in this case, wave spectral density) as

where ______________________ is the RBF shape parameter, and ‘N’ is the total number of nodes. The wave action balance equation is then projected on sphere as

and a unique n x n stencil (n≪N) is constructed for each node. For deep water and without currents or source terms, the numerical scheme (with hyperviscosity) becomes

where cg is the group velocity, ‘!’ represents element multiplication, and DN is a sparse N x N differentiation matrix. This can be solved using a parallelized explicit time-stepping method.

VII. Unstructured Node PrototypeThe previously mentioned disadvantages (pole singularity and shrinking grid cells) are an artifact of representing 3D spherical data on a 2D plane and a decreasing time step due to the CFL condition. One possibility being explored is the use of unstructured nodes on a 3D sphere (both spatial and wavenumber) to remove the singularity and spread the computational costs equally.

Radial basis functions (RBFs) use an arbitrary node layout and are well-suited to solving convective PDEs on a sphere due to their algorithmic simplicity and spectral accuracy. However, the method is computationally expensive when scaled to a large number of 'N' nodes. To circumvent this, the RBF generated finite-difference (RBF-FD) method uses large stencils to generate differential weights to re-produce RBFs (rather than polynomials). This meth-od scales as O(N) per time step and is parallelizable.

Figure 7: Example of an unstructured node layout to solve the wave action balance equation on an aquaplanet.

∂tN +P∇k Ω ·P∇x N +P∇x Ω ·P∇k N = P(Source Terms)

N (α, t)t=t0

=N

i=1

λi φ(α−αi), for φ(r) = e−(εr)2 and ε, r ∈ R+,

∂tN = −cgx DxNN − cgy Dy

NN − cgz DzNN

Adrean WebbDept. of Applied Math, [email protected]

Global Stokes Drift andClimate Wave Modeling

Baylor Fox-KemperDept. of Atm. and Oceanic Sciences, CIRES

[email protected]

The Stokes drift velocity - the Lagrangian-average fluid velocity induced from wave action - is an important vector component that appears often in wave-averaged dynamics, such as Langmuir turbulence. However, accuracy and data coverage remain challenges in estimating Stokes drift globally and recent research (Hanley et al., 2010, Webb & Fox-Kemper, 2011) has shown that using atmospheric data and the assumption of wave equilibration is not trustworthy, as often the wave state is dominated by developing or remotely-generated swell conditions.

Figure 1: Example of the global variability of Stokes drift (created using WWATCH III and CORE2 winds).

II. Importance for Climate ResearchStokes drift is of interest in climate research due to its dominant role in determining the strength of Langmuir turbulence - surface mixing due to the interaction of wind and waves. Preliminary work (Fox-Kemper et al, 2011) to include this unresolved turbulence in climate models, has shown the potential to correct a well-known, shallow mixed-layer bias in the Southern Ocean.

Figure 2: (a) Example of Langmuir turbulence (Deep Water Horizon spill), (b) Southern Ocean mixed-layer depth results from a preliminary parametrization of Langmuir turbulence (NCAR CCSM 3.5)

npr.org

VI. Need for a Climate Wave ModelUnfortunately, all conventional, structured latitude-longitude-grid wave-models have a disadvan-tage in that they are singular at the poles and the grid cell sizes shrink appreciably with higher latitudes. This is a problem for climate simulations since the Arctic may be ice-free in the near future and performance of the model will deteriorate as the northern boundary is moved higher. Overcoming these problems and improving performance for long model runs (without compromising climate-scale statistics), remains a challenge.

V. Coupling Wave and Circulation ModelsFor climate research purposes, it is advantageous to use a prognostic 2D wave field - not only to calculate Langmuir turbulence but also to improve atmospheric and oceanic fluxes. As such, NCAR is currently adding a modified form of WWATCH III (v3.14) to the development code of the Community Earth System Model (CESM). The wave module component will be linked to the atmosphere-ocean coupler on 30 minute intervals. Since it is not uncommon for climate runs to span several hundred years, a high model throughput of 30 or more simulated years per computational day is ideal and early benchmarking shows that this is achievable if a coarsened grid of 3.2 x 4 deg is used with a northern and southern latitude cutoff of 70 deg.

Figure 6: WWATCH III grid performance with early benchmarking targets on two different machines: (a) NASA Pleaides, (b) NCAR Bluefire. Performance is measured in number of simulated years per day of running.

30 simulated years per day or greater is necessary to couple with ECSM

α = (x, y, z, kx, ky, kz), ε

VIII. ConclusionIn computations where the accuracy of Stokes drift is important, it is essential to use either the full 2D wave spectra or 1D spectra with an empirically-derived spreading function. Currently, WWATCH III is being coupled to the NCAR CESM to add Langmuir turbulence and improve air-sea fluxes. In addition, a prototype climate wave model is being developed for long model runs and ice-free conditions in the Arctic. While there are still major challenges to overcome (such as boundaries and parallelization), this unstructured node approach shows promise.

III. Leading-order Stokes DriftThe Stokes drift can be defined as the mean temporal and spatial difference between the Eulerian and Lagrangian velocities for a finite period and length scale. Using a series expansion and a linear wave decomposition, the leading-order Stokes drift can be rewritten in spectral density form as

Notice that the leading order Stokes drift is a vector quantity whose magnitude depends both on the directional components of the wave field and the directional spread of wave energy for each component.

To illustrate how different directional wave components might affect the magnitude of Stokes drift, consider a pair of monochromatic waves traveling with an incident angle !’ from the y-axis (Fig. 3). Then the magnitude of Stokes drift will depend on sin !’.

Figure 3: An example of how the directional components of a wave field can affect the magnitude of the Stokes drift.

uS ≈ 16π3

g

0

π

−π

cos θ, sin θ, 0

f3 Sfθ(f, θ) e

8π2f2

g zdθ df

!’

IV. One-dimensional Stokes Drift

Commonly, a unidirectional wave approxi-mation (|H|=1) is used to simplify the 1D Stokes drift. However, this tends to over-estimate the full leading-order Stokes drift by 33% (in a 3rd generation wave model) and is not recommended (Webb & Fox-Kemper, 2011).

A better approximation can be found by using an empirical ly-determined frequency-dependent spread function for "f. To illustrate, a ‘directional loss’ is plotted in Fig. 5 using a spread function derived by Donelan et al. (1985).

0.95,0.93

0.56,0.78 1.6,0.780 0.5 1 1.5 20.7

0.8

0.9

1

f fpHff p

Often only 1D spectral data, or even just spectral moments, are available and this poses a challenge to accurately estimate Stokes drift (see Fig. 4 for example).

16.4. OCEAN-WAVE SPECTRA 285

0

10

30

50

60

80

90

110

100

70

40

20

0 0.05 0.10 0.15 0.20 0.25 0.30 0.35Frequency (Hz)

Wa

ve S

pe

ctra

l De

nsi

ty (

m2

/ H

z)

20.6 m/s

18 m/s

15.4 m/s

12.9 m/s

10.3 m/s

Figure 16.7 Wave spectra of a fully developed sea for di!erentwind speeds according to Moskowitz (1964).

Pierson-Moskowitz Spectrum Various idealized spectra are used to answerthe question in oceanography and ocean engineering. Perhaps the simplest isthat proposed by Pierson and Moskowitz (1964). They assumed that if thewind blew steadily for a long time over a large area, the waves would come intoequilibrium with the wind. This is the concept of a fully developed sea. Here, a“long time” is roughly ten-thousand wave periods, and a “large area” is roughlyfive-thousand wave lengths on a side.

To obtain a spectrum of a fully developed sea, they used measurements ofwaves made by accelerometers on British weather ships in the north Atlantic.First, they selected wave data for times when the wind had blown steadily forlong times over large areas of the north Atlantic. Then they calculated the wavespectra for various wind speeds (figure 16.7), and they found that the function

S(!) ="g2

!5exp

!

!#"!0

!

#4$

(16.28)

was a good fit to the observed spectra, where ! = 2$f , f is the wave frequencyin Hertz, " = 8.1 " 10!3, # = 0.74 , !0 = g/U19.5 and U19.5 is the wind speedat a height of 19.5 m above the sea surface, the height of the anemometers onthe weather ships used by Pierson and Moskowitz (1964).

For most airflow over the sea the atmospheric boundary layer has nearly

As illustrated previously, 1D estimates will tend to overestimate the full leading-order Stokes drift when there is more than one directional component. However, the error due to wave spreading can be minimized by using an empirical spreading function. Consider splitting the 2D spectral density as

where "f is the ‘directional distribution’. Then a directional loss H due to wave energy being directed along other directions other than the dominant direction, can be defined as

and the 1D Stokes drift approximation can be written as

0

π

−πSfθ(f, θ) dθdf =

0

π

−πφf (f, θ)Sf (f) dθdf =

0Sf (f) df,

H(f) =

π

−π

cos θ, sin θ, 0

φf (f, θ) dθ

Figure 4: Examples of wave spectra: (a) 2D spectra generated by WWATCH III, (b) 1D Pierson and Moskowitz observational spectra (Stewart, Intro. to Physical Oceanography).

Figure 5: An example of directional spread loss using an empirically-determined spread function from Donelan et al. (1985).

AcknowledgementsThanks to Natasha Flyer at NCAR for advice on implementing the RBF-FD method. Work supported by NASA NNX09AF38G, NSF 0934737, CIRES, UCAR, and CU-Boulder.

ReferencesFornberg, B., Lehto, E., 2011. Stabilization of RBF-generated finite difference methods for convective PDEs, Journal of Computational Physics,

Volume 230, Issue 6.Donelan, M.A., Hamilton, J., Hui, W.H., 1985. Directional spectra of wind-generated waves. Philosophical Transactions of the Royal Society of

London Series A - Mathematical Physical and Enineering Sciences 315, 509-562.Fox-Kemper, B., Webb, A., Baldwin-Stevens, E., Danabasoglu, G., Hamlington, B., Large, WG., Peacock, S., in preparation. Global climate model

sensitivity to estimated Langmuir Mixing. Ocean Modelling.Hanley, K.E., Belcher, S.E., Sullivan, P.P., 2010. A global climatology of wind–wave interaction. Journal of Physical Oceanography 40, 1263–1282.Flyer, N., Wright, G., 2009. A Radial Basis Function Method for the Shallow Water Equations on a Sphere. Proceedings of the Royal Society A -

Mathematical Physical and Engineering Sciences.Webb, A., Fox-Kemper, B., 2011. Wave spectral moments and Stokes drift estimation. Ocean Modelling.

I. Introduction

us ≈ 16π3

g

0H(f) f3 Sf (f) e

8π2f2

g z df

Wednesday, November 30, 2011