06.02.2014 Views

NSF proposal - Physical Oceanography Research Division

NSF proposal - Physical Oceanography Research Division

NSF proposal - Physical Oceanography Research Division

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Project Summary<br />

Eddy diffusivities in the Southern Ocean from an eddying model<br />

Alexa Griesel, Sarah T. Gille, Julie L. McClean<br />

Summary. Realistic depiction of Southern Ocean eddy dynamics is critical in models<br />

used for climate prediction. Mixing generated by mesoscale eddies is believed to play<br />

an important role in the transfer of water masses and tracers across the Antarctic Circumpolar<br />

Current (ACC). In turn, these water mass transfers are thought to control the<br />

global overturning circulation and inter-ocean exchange. The central goals of this <strong>proposal</strong><br />

are to (1) quantify Lagrangian isopycnal eddy diffusivities in the Southern Ocean<br />

and to characterize their horizontal and vertical distributions, and (2) to assess their ability<br />

to parameterize isopycnal eddy tracer fluxes across the ACC. We will address these<br />

questions using the 0.1 ◦ , 42-level global configuration of the Los Alamos National Laboratory<br />

(LANL) Parallel Ocean Program (POP).<br />

Intellectual Merit. Large uncertainty exists in the representation of Southern Ocean eddy<br />

mixing processes in standard climate scale ocean models. The effects of eddies are commonly<br />

parameterized using eddy diffusion, and climate scale models are very sensitive to<br />

the magnitude of the eddy diffusion coefficients in the Southern Ocean. Whether the sensitivity<br />

of the ocean circulation to surface forcing is altered when using constant eddy diffusion<br />

coefficients rather than variable eddy diffusion coefficients remains unclear, and,<br />

more fundamentally, if local, or even zonally integrated tracer transports are simulated<br />

adequately by the eddy diffusion model. A related fundamental issue is to determine the<br />

horizontal and vertical distributions of the eddy diffusion coefficients. Different methods<br />

have resulted in conflicting magnitudes and spatial distributions in the Southern Ocean.<br />

Lagrangian floats provide a way to test the applicability of the diffusion model. We propose<br />

to deploy O(10,000) numerical floats in POP, at various locations within and outside<br />

the ACC and at a range of depth levels to estimate Lagrangian diffusivities and their spatial<br />

distributions. The skill of these Lagrangian diffusivities in parameterizing zonally<br />

integrated as well as local eddy tracer transport in the ACC will be assessed. This proposed<br />

work will contribute to ongoing efforts to reconcile conflicting diffusivity estimates<br />

for the Southern Ocean and will ultimately allow us to assess how much the spatial structure<br />

of the diffusivities matters in simulating Southern Ocean eddy fluxes accurately.<br />

.<br />

Broader Impacts. The proposed work will contribute to early career development of a<br />

postdoctoral researcher. The project will explore different ways to parameterize mixing<br />

coefficients, which will be testable in predictive climate models. A detailed assessment<br />

of Lagrangian methods and their relation to eddy fluxes and diffusivity estimates from<br />

other methods will benefit efforts to analyze data from floats deployed all over the ocean.<br />

In particular, the methods investigated in this study will contribute to the interpretation<br />

of in situ data returned by the 180 RAFOS floats that are being deployed as part of the<br />

Diapycnal and Isopycnal Mixing Experiment in the Southern Ocean (DIMES).<br />

B–1


Results from Prior <strong>NSF</strong> Support<br />

A. Griesel: no prior <strong>NSF</strong> support.<br />

S.T. Gille: Collaborative <strong>Research</strong>: DIMES, Diapycnal and Isopycnal Mixing in the<br />

Southern Ocean. OCE-0622740 (Sarah Gille); 7/1/07-6/30/12, $555,192. DIMES is a<br />

UK/US effort to measure diapycnal mixing processes and along-isopycnal stirring along<br />

the steeply tilted isopycnals of the Southern Ocean. Fieldwork successfully began in January<br />

2009. Gille et al. (2007) summarize the DIMES objectives. At Scripps, preliminary<br />

work in preparation for the field measurements has focused on analysis of Lagrangian<br />

floats released in the Parallel Ocean Program (Griesel et al., 009a,b). We have also carried<br />

out work to evaluate the vertical resolution of XCTD data (Gille et al., 2009).<br />

J. McClean: Title: Mesoscale Variability and Processes in an Eddy-Resolving Global<br />

POP Simulation. OCE-0221781/0549225: 09/01/2002-08/31/2008, $391,168. The mean<br />

and variability of the upper ocean circulation of global 0.1 ◦ POP was compared with data<br />

from altimetry, current meters, time series of temperature profiles, and published results<br />

of volume transports and meridional heat transports (Maltrud and McClean, 2005; Mc-<br />

Clean et al., 2006, 2008). A 0.1 ◦ North Atlantic configuration of POP was compared with<br />

tide gauges and altimetry using wavelet transforms (Tokmakian and McClean, 2003).<br />

Interannual processes in the Indo-Pacific as depicted by POP were examined (McClean<br />

et al., 2005; Prasad and McClean, 2004). Studies of mesoscale properties were conducted<br />

in the Agulhas Retroflection (Byrne and McClean, 2008) and off Japan in the Kuroshio<br />

Extension System Study (KESS) data collection region (Rainville et al., 2007). Model eddy<br />

heat fluxes and the appropriateness of diffusive parameterizations were examined in the<br />

Southern Ocean (Griesel et al., 009a,b). Model output was disseminated broadly throughout<br />

the oceanographic community.<br />

PROJECT DESCRIPTION<br />

1 Introduction<br />

Eddy mixing plays a unique role in the horizontal and vertical transfer of tracers across<br />

the Antarctic Circumpolar Current (ACC), and it is thought to control the global circulation<br />

of heat, carbon and other climatically important tracers through its impact on<br />

overturning rate and inter-ocean exchanges (Johnson and Bryden, 1989; Marshall et al.,<br />

1993; Döös and Webb, 1994; Speer et al., 2000; Bryden and Cunningham, 2003; Marshall<br />

and Radko, 2003). The Southern Ocean has undergone significant climate change (Gille,<br />

2002, 2008; Böning et al., 2008), and eddy fluxes may significantly contribute to the observed<br />

warming in the Southern Ocean (e.g. Hogg et al., 2008). The classical paradigm<br />

of the zonal mean circulation in the Southern Ocean is one where wind driven Ekman<br />

transport that tends to steepen isopycnals is approximately balanced by an eddy-driven<br />

circulation that tends to flatten isopycnals. This leads to a residual circulation that governs<br />

net exchange of tracers (Karsten et al., 2002; Karsten and Marshall, 2002; Marshall<br />

and Radko, 2003). However, the degree of cancellation is unclear and large zonal varia-<br />

D–1


tions in residual circulation may exist (Treguier et al., 2007). The characterizaton of the<br />

three-dimensional distribution of eddy-driven mixing in the ACC is therefore crucial and<br />

remains highly uncertain.<br />

Numerical models typically represent eddy-driven mixing as a downgradient transfer<br />

along isopycnals at a rate proportional to a diffusivity coefficient (Gent and McWilliams,<br />

1990; Griffies, 1998). Coarse resolution models are very sensitive to the magnitude, and<br />

the horizontal and vertical distributions of the eddy mixing coefficients (Kamenkovich<br />

and Sarachik, 2004; Griesel and Morales Maqueda, 2006; Danabasoglu and Marshall,<br />

2007). Moreover, the transient behavior of a model with explicit eddies can be quite different<br />

from the behavior of one in which eddy effects are parameterized with a constant<br />

diffusivity (Hallberg and Gnanadesikan, 2006).<br />

Two fundamental questions drive the work proposed here:<br />

(1) Can eddy diffusion provide an appropriate parameterization of eddy fluxes?<br />

(2) What are the magnitudes and spatial distributions of the diffusion coefficients?<br />

We propose to address these questions in the framework of the 0.1 ◦ , 42-level global configuration<br />

of the Los Alamos National Laboratory (LANL) Parallel Ocean Program (POP)<br />

(Maltrud et al., 2008). We plan to evaluate how well eddy parameterizations can emulate<br />

the complex spatio-temporal effects of eddies in the Southern Ocean of POP by<br />

diagnosing Eulerian eddy tracer fluxes and their relation to Eulerian mean gradients, and<br />

also by means of Lagrangian statistics. To examine Lagrangian statistics, we will deploy<br />

O(10,000) numerical floats at locations both inside and outside the ACC and at a range of<br />

depth levels.<br />

Lagrangian floats provide a way to test the applicability of the eddy diffusion model<br />

(see LaCasce, 2008, for an overview). Taylor (1921) showed that Lagrangian particles<br />

in isotropic and homogeneous turbulence spread diffusively, with a constant diffusivity<br />

calculated from the integral of the velocity autocorrelation. Davis (1987, 1991) refined<br />

the theoretical framework to allow both the mean flow and the diffusivity to be spatially<br />

variable, and to allow the theory to be applied to ocean float data. Preliminary results<br />

with a limited number of floats in the POP model suggest that, under certain conditions,<br />

a diffusive limit can be reached in the ACC, providing a basis for the proposed work<br />

(Griesel et al., 009b). Only a few investigations have studied Lagrangian statistics in the<br />

ACC (e.g. Sallée et al., 2008), and of the work we have seen, only our own preliminary<br />

study (Griesel et al., 009b) has studied their depth dependence, albeit with limited spatial<br />

resolution. Davis (1987, 1991) theorized on the mechanisms by which the Lagrangian<br />

single-particle statistics are related to Eulerian eddy tracer fluxes through an elaboration<br />

of the familiar eddy diffusion model in an eddy field with finite scales. No previous study<br />

that we have identified has investigated in detail the extent to which Lagrangian statistics<br />

are consistent with Davis (1987, 1991)’s model in terms of emulating the Eulerian eddy<br />

fluxes in a realistic configuration of the Southern Ocean.<br />

D–2


Figure 1: Vertical profiles of eddy<br />

kinetic energy ( 1 2 (u′2 + v ′2 )) for<br />

AUSSAF data compared with profiles<br />

from other sites in the ACC:<br />

southeast of New Zealand and<br />

Drake Passage (solid lines) and<br />

from western boundary current<br />

systems: Agulhas, Gulf Stream<br />

and Kuroshio Extension (dashed<br />

lines), from Phillips and Rintoul<br />

(2000).<br />

2 Spatial distribution of eddy mixing in the Southern Ocean<br />

In the absence of direct measurements of eddy mixing across the ACC, a number of studies<br />

have attempted to deduce horizontal and vertical distributions of eddy diffusivities.<br />

These studies have produced conflicting results, although, as we will explain below, they<br />

are consistent with different hypotheses for eddy diffusivity distributions. On the one<br />

hand, eddy diffusivities are hypothesized to be determined by eddy kinetic energy (EKE)<br />

(e.g. Keffer and Holloway, 1988; Stammer, 1998). A recent study found variations in eddy<br />

stirring and mixing rates using finite-time Lyapunov exponents to be closely related to<br />

variations in mesoscale activity, or √ EKE (Waugh and Abraham, 2008). Diffusivities inferred<br />

from Lagrangian floats are also often highly correlated with eddy kinetic energy<br />

(Krauss and Böning, 1987; Lumpkin et al., 2001; Zhurbas and Oh, 2003; Sallée et al., 2008),<br />

resulting in local maxima of eddy diffusivity in the core of the ACC (Sallée et al., 2008).<br />

On the other hand, strong currents can also act as mixing barriers (e.g. Bower et al., 1985),<br />

leading to small effective eddy diffusivities in the core of the ACC (e.g. Marshall et al.,<br />

2006; Naveira Garabato et al., 2009). Naveira Garabato et al. (2009) employed mixing<br />

length arguments, and found that it is the spatial distribution of the eddy length scales,<br />

rather than the eddy velocities or EKE that determines the spatial distributions of the<br />

eddy diffusivities. At any given geographic location, the magnitude of the diffusivity<br />

depends on the relative importance of the strength of the mean flow and EKE respectively<br />

(Shuckburgh et al., 2008a,b). Griesel et al. (009b) reconciled some of the conflicting<br />

horizontal near-surface eddy diffusivity distributions by pointing out that dominant correlations<br />

of Lagrangian diffusivities with EKE may have been related to not using long<br />

enough time lags in the integral of the Lagrangian velocity autocovariance. However, the<br />

float deployments used by Griesel et al. (009b) were limited, both in number and spatial<br />

distribution, and did not resolve the ACC and the region north of the ACC sufficiently to<br />

D–3


Figure 2: Depth dependence of alongstreamline<br />

averaged effective eddy diffusivities,<br />

estimated from tracers advected with<br />

the velocity of the Southern Ocean State<br />

estimate (SOSE) from Abernathy et al.<br />

(2008).<br />

distinguish regimes with strong and weak mean flow.<br />

There is also a lack of consensus on the depth dependence of the eddy diffusivities.<br />

Diffusivities are predicted to decrease with depth when the depth dependence is dominated<br />

by the decrease of eddy velocity or EKE with depth (see Figure 1). Such a decrease<br />

has been found by Ferreira et al. (2005); Eden (2006); Olbers and Visbeck (2005)<br />

and Griesel et al. (009a). Conversely, critical layer theory predicts that the diffusivity<br />

should have a maximum at the depth where the speed of the mean flow is comparable<br />

to the phase speed of Rossby waves (Green, 1970; Smith and Marshall, 2009). Indeed,<br />

enhanced diffusivities at depth have been diagnosed in eddying models of varying complexity<br />

by Cerovecki et al. (2009); Smith and Marshall (2009) and Abernathy et al. (2008)<br />

(see Figure 2), using the Nakamura (1996) method, and from flux/gradient relationships<br />

(Treguier, 1999; Smith and Marshall, 2009). Of the studies listed here, only our investigation<br />

(Griesel et al., 009b) used Lagrangian methods to investigate depth dependence. We<br />

found that depth invariant Lagrangian diffusivity in the core of the ACC jet arose from a<br />

combination of eddy velocities decreasing and Lagrangian eddy length scales increasing<br />

with depth.<br />

Even though eddy mixing can be diagnosed with different methods, the fundamental<br />

issue remains whether these eddy diffusion coefficients actually have sufficient skill to parameterize<br />

eddy transports in a downgradient parameterization. Challenges arise when<br />

attempting to correlate downgradient parameterizations with the eddy fluxes. One reason<br />

for low correlations of the eddy fluxes with their parameterizations and for unphysically<br />

large positive and negative values of the diffusivities is the presence of strong rotational<br />

components (Lau and Wallace, 1979; Marshall and Shutts, 1981, henceforth MS81).<br />

These rotational components of the eddy flux do not contribute to the local heat budget,<br />

since they transport as much heat into a region as out of it (Jayne and Marotzke, 2002).<br />

In previous work with POP Griesel et al. (009a) showed that the rotational component of<br />

the eddy heat flux dominates over all spatial scales. We also pointed out the challenges<br />

of finding eddy diffusion coefficients that are physically meaningful, smoothly varying<br />

and most importantly, that still capture most of the eddy flux in a downgradient parameterization.<br />

On the other hand, Griesel et al. (009b) have shown that the diffusive limit<br />

can be reached over lengthscales of 10 ◦ longitude with Lagrangian floats. One aim of<br />

D–4


the proposed work is to reconcile these two findings. Significantly more numerical float<br />

deployments than the O(100) deployments used in Griesel et al. (009b) are needed in order<br />

to sufficiently resolve spatial variability of Lagrangian diffusivities. Since Lagrangian<br />

floats are widely used and are an invaluable tool to assess eddy mixing from observations,<br />

it is important to be able to reconcile the previously reported estimates of eddy<br />

diffusivities that used different methods, as well as compare the Eulerian analysis with<br />

the Lagrangian eddy statistics. This will allow better interpretation of Lagrangian data<br />

that emerge from field programs.<br />

The Diapycnal and Isopycnal Mixing Experiment in the Southern Ocean (DIMES) will<br />

deploy a total of 180 RAFOS floats into the Southern Ocean. While DIMES will deliver a<br />

wealth of data on eddy mixing in the Southern Ocean, it is crucial to be able to interprete<br />

these properly. It is not feasible to deploy O(10,000) in the real ocean, and we will be able<br />

to develop the strategies for achieving robust results with the more limited data from the<br />

observational program.<br />

3 Objectives<br />

The first overarching goal of the proposed work is to characterize magnitudes and horizontal<br />

and vertical distributions of the Lagrangian eddy diffusivities in the Southern<br />

Ocean and to evaluate whether these can be reconciled with existing hypotheses. The<br />

hypotheses are that eddy diffusivities are to first order related to the spatial distributions<br />

of eddy kinetic energy, which would result in diffusivities being elevated at the surface<br />

and decreasing with depth in the core of the ACC. The alternative hypothesis is that eddy<br />

mixing is to first order determined by the relative strength of the mean flow and phase<br />

speed of Rossby waves, leading to the ACC jet acting as a mixing barrier at the surface,<br />

and diffusivities increasing in the core of the ACC jet.<br />

The second overarching goal is to investigate how well these diffusivities parameterize<br />

the integrated and local eddy tracer transports across the ACC. We propose to<br />

deploy a large number of numerical floats (O(10,000)) in POP, to quantify Lagrangian<br />

diffusivities and their spatial distributions, and to evaluate the relation of Eulerian eddy<br />

tracer fluxes and mean tracer gradients to assess the hypothesis that Lagrangian floats<br />

can be used to quantify eddy mixing and its spatial variation in the Southern Ocean.<br />

4 Approach and Methodology<br />

4.1 Objective 1: Characterizing the effects of eddies as a diffusive process<br />

using Lagrangian statistics<br />

4.1.1 Background<br />

Davis (1991) showed that the evolution of the mean tracer concentration in inhomogeneous<br />

flows with finite scale eddies depends only on the initial condition of the mean<br />

and source fields and the single particle statistics. The single particle Lagrangian diffu-<br />

D–5


sivity as a function of time lag is related to the Eulerian eddy fluxes and mean tracer<br />

gradients through the elaborated flux versus gradient law,<br />

〈u ′ i (x, t)Θ′ (x, t)〉 = lim τ−→∞ −<br />

∫ t<br />

0<br />

dτ∂ τ κ ij (x, τ)∂ xj Θ(x, t − τ) (1)<br />

Equation (1) specifies the flux of a (passive) tracer from the recent history of its mean<br />

concentration. The driving force for the eddy flux at time t is the sum of previous Θ gradients<br />

at times t − τ, weighted by the time derivative of κ at time range τ. κ will be defined<br />

below. This is an elaboration of the familiar advection-diffusion equation, where the elaboration<br />

reflects the finite scales of the eddies. If τ goes to infinity, or rather, is greater than<br />

some time T, that reflects the finite scales of the dispersing eddies, κ(x, t) approaches a<br />

constant value κ ∞ and the eddy flux is approximated by the typical advection-diffusion<br />

equation,<br />

〈u ′ i (x, t)Θ′ (x, t)〉 = −κ ∞ ij (x)∂ x j<br />

Θ(x, t) (2)<br />

The diffusivity κ ∞ is determined by the integral of the Lagrangian autocovariance<br />

function:<br />

κ ij (x, τ) =<br />

∫ 0<br />

−τ<br />

d ˜τ〈u ′ i (t 0|x, t 0 ) u ′ j (t 0 + ˜τ|x, t 0 )〉 L , (3)<br />

κ ∞ ij = lim τ−→∞ κ ij (x, τ), (4)<br />

where u ′ i (t 0 + τ|x, t 0 ) denotes the residual velocity of a particle at time t 0 + τ passing<br />

through x at time t 0 . For stationary and homogeneous statistics, the diffusivity is equal to<br />

half the time rate of change of the dispersion 〈d ′2 〉<br />

κ ij (x, τ) = 1 2 d t〈d ′ i (τ|x, t 0)d ′ j (τ|x, t 0)〉 L , (5)<br />

where d ′ (τ|x, t 0 ) is the displacement at time τ from its starting position, with the mean<br />

displacement removed. There are a number of challenges when computing Lagrangian<br />

diffusivities from eddying models, or oceanic data. To isolate the eddy velocities from the<br />

float velocities, a mean needs to be subtracted. As Figure 3a shows, the time-averaged Eulerian<br />

mean velocities are highly inhomogeneous in space. As shown by Oh et al. (2000),<br />

in a shear flow the along-stream diffusivity is affected by the shear dispersion. Therefore,<br />

if we remove a spatially-uniform average from each float velocity, we are left with a residual<br />

due to the mean shear that will imply dispersion and may dominate the diffusivity<br />

estimate (Bauer et al., 1998). Griesel et al. (009b) achieved good convergence properties<br />

of the Lagrangian diffusivities when they subtracted this Eulerian mean, thereby minimizing<br />

the influence of shear dispersion (Figure 3b,c). In an ocean with inhomogeneous,<br />

strong mean flows and with an eddy field of varying horizontal scale, the challenge is to<br />

find the asymptotic behavior of the velocity autocorrelation, which requires sufficiently<br />

long time lags (e.g. Bauer et al., 1998; Veneziani et al., 2004; Davis, 2005). The challenge<br />

arises because of the presence of strong rotational components in the eddy field, showing<br />

up in vortices or meanders of the Lagrangian trajectories that lead to oscillations in the autocovariances<br />

(Figure 4) as discussed by Berloff and McWilliams (2002a,b) and Veneziani<br />

D–6


a)<br />

⊥ dispersion (km 2 )<br />

2 x 1011<br />

b) ⊥ 300m<br />

1<br />

0<br />

0 250 500 750 1000<br />

time lag (days)<br />

|| dispersion (km 2 )<br />

3 x 1012 c) || 300m<br />

2<br />

1<br />

0<br />

0 250 500 750 1000<br />

time lag (days)<br />

Figure 3: a) Trajectories of 160 floats deployed at a depth of 300 m around latitude 60 ◦ S and<br />

longitude 180 ◦ W in one-quarter degree initial spacing from Griesel et al. (009b) after they have<br />

travelled for 1045 days (about 3 years). In color is the magnitude of the Eulerian 1998-1999 mean<br />

velocity interpolated to each float point. Lower panels show the corresponding long-time (1045<br />

days) mean square displacement for these trajectories in the cross-stream (b) and along-stream<br />

directions (c). The dispersion was calculated after subtraction of the spatially highly non-uniform<br />

Eulerian mean velocity in (a) from every float velocity. Shaded are 2*standard deviations from<br />

bootstrapping by subsampling the floats for each deployment with replacement. The dispersion<br />

is well approximated by the constant diffusivity at a time lag of 100 days (solid lines in a and b,<br />

where the dashed lines are the 95% confidence intervals).<br />

et al. (2005). Questions remain as to whether this is a transient behavior, or whether the<br />

diffusive limit can still be reached after long enough times lags (see Figure 3b,c), as suggested<br />

by Griesel et al. (009b).<br />

The Lagrangian diffusivity of Taylor can be expressed as the product of the eddy kinetic<br />

energy and an eddy decorrelation time scale T e , or, alternatively, with an eddy length<br />

scale L e = √ (EKE)T e . This eddy length scale can be interpreted as the mean displacement<br />

induced by an eddy. It is unclear how this eddy length scale is related to the typical<br />

eddy mixing length of for example Armi and Stommel (1983). Ferrari and Nikurashin<br />

(2009) and Naveira Garabato et al. (2009) hypothesize that the true mixing length of eddies<br />

is proportional to the Lagrangian eddy length scale divided by 1 + (U m − c) 2 , where<br />

U m is the strength of the mean flow and c is the phase speed of Rossby waves. In that<br />

sense, mixing length itself is dependent on eddy kinetic energy, and the eddy mixing<br />

length equals the Taylor eddy length scale, but reduced by the strength of the mean flow.<br />

This is in contrast to the assumption that in the presence of a mean flow the Taylor diffusivity<br />

by itself, as refined by Davis (1987, 1991), is able to quantify eddy mixing.<br />

D–7


cross−stream direction<br />

latitude<br />

300<br />

200<br />

100<br />

0<br />

−100<br />

−52<br />

−54<br />

−56<br />

−58<br />

−60<br />

along−stream direction<br />

−62<br />

180 190 200 210 220<br />

longitude<br />

3000<br />

a) b)<br />

c)<br />

κ m 2 s −1 )<br />

κ (m 2 s −1 )<br />

2000<br />

1000<br />

0<br />

−1000<br />

3000<br />

2000<br />

1000<br />

10 20 30<br />

tlag (days)<br />

d)<br />

0<br />

0 20 40 60 80<br />

tlag (days)<br />

Figure 4: Circling and meandering trajectories<br />

(a) do not lead to any net cross-stream<br />

dispersion and the diffusivity as a function<br />

of time lag oscillates before it converges to<br />

the correct value of zero diffusivity (b). Real<br />

POP trajectories from Griesel et al. (009b) (c)<br />

show this oscillating behavior (d). Enough<br />

float trajectories such that the oscillations<br />

may cancel each other are needed as well as<br />

a long enough time lag to reach κ ∞<br />

4.1.2 Float Deployments in POP<br />

The Los Alamos National Laboratory (LANL) Parallel Ocean Program (POP) has been<br />

configured on a 0.1 ◦ , 42-level global tripole grid, which is Mercator south of 28.5 ◦ N (Maltrud<br />

et al., 2008). The POP model has one of the highest horizontal resolutions used<br />

in global ocean modeling applications, leading to horizontal resolutions between 4 and<br />

9 km for the Southern Ocean. POP is eddy resolving in the sense that between about<br />

50 ◦ S and 50 ◦ N it resolves the first baroclinic Rossby radius, which is smaller than typical<br />

eddy length scales (Smith et al., 2000). To date the simulation has been run for 100 years<br />

and is forced with the monthly normal year atmospheric fluxes constructed by Large and<br />

Yeager (2004) for the Common Ocean-Ice Reference Experiments (CORE). Normal year<br />

forcing consists of single annual cycles of all atmospheric data needed to force an ocean<br />

model and is representative of climatological conditions over 43 years of interannually<br />

varying atmospheric state. It has been constructed to retain synoptic variability (i.e. atmospheric<br />

storms) with a seamless transition from 31 December to 1 January (Large and<br />

Yeager, 2004). The forcing uses the National Center for Environmental Prediction/ National<br />

Center for Atmospheric <strong>Research</strong> (NCEP/NCAR) reanalysis corrected with scatterometer<br />

data. A marked improvement was seen in the pathways of the Agulhas eddies,<br />

relative to the global 0.1 ◦ , 40-level standard z-level dipole simulation of Maltrud and Mc-<br />

Clean (2005). Here we propose to conduct a 5-10 year global tripole 0.1 ◦ POP simulation<br />

forced with interannually-varying synoptic CORE forcing for the more recent years 1997-<br />

2007 initialized from this spun-up ocean state. The exact duration of the run will depend<br />

on the throughput of the <strong>NSF</strong> Terascale high performance computing platforms available<br />

to us. We will, however, be able to archive 2-3 years of daily state variables and flux terms,<br />

and otherwise monthly means.<br />

To analyze Lagrangian motions, it is crucial that we are able to advect the POP floats<br />

as part of the model runs, using the time-varying three-dimensional velocity field. The<br />

position of the floats will be evaluated with a fourth-order Runge Kutta scheme at every<br />

model time step of 6 minutes. This is clearly an advantage over previous studies that<br />

used multiple-day averages of velocities to advect floats. As shown by (Griesel et al.,<br />

D–8


009b), floats advected by the three-dimensional model field stay on isopycnal surfaces<br />

reasonably well, at least away from the oceanic surface and mixed layers. However, small<br />

deviations from isopycnals are evidence of small diapycnal mixing rates that can be crucial<br />

for Southern Ocean dynamics. Since there is great interest in understanding how<br />

three-dimensional motions may differ from isopycnal motions, we will also explore ways<br />

to correct the motion of the floats to force them to stay on an isopycnal. Additionally,<br />

there is the potential to investigate sensitivities of float trajectories to different advection<br />

schemes as well as model parameters.<br />

We propose to deploy a total of O(10,000) numerical floats in the following ways:<br />

(1) The depth dependence and horizontal variability of Lagrangian diffusivities has been<br />

investigated by Griesel et al. (009b), but spatial coverage of the floats was limited to the<br />

area within the ACC, and due to the limited number of floats, bins with significant realizations<br />

were sparse. To improve upon the numerical float deployments in POP, to<br />

achieve more spatial coverage within the ACC and also to cover the area north of the<br />

ACC, we propose to deploy floats at 4 different longitudes with a quarter degree initial<br />

spacing, and at latitudes corresponding to south of the Polar Frontal Zone, between the<br />

Polar Frontal Zone and Subantractic Front, and north of the Subantarctic Front. ”Point”<br />

deployments in this way will make it possible to track the spreading of a group of floats<br />

along one specific isopycnal interval. The mean position of the fronts will be estimated<br />

with time mean barotropic streamlines from the model. To sample with depth, the floats<br />

will be deployed at 4 different depth levels.<br />

(2) One deployment will be configured to match that of the DIMES experimental deployment<br />

scheme (around 105 ◦ W, 55 ◦ -60 ◦ S, and at two different depth levels, or isopycnal<br />

surfaces), to determine how well the limited number of floats deployed as part of DIMES<br />

can reproduce the magnitude and spatial variability of the extended deployments, and to<br />

help improve analysis methods for the real floats.<br />

(3) To assess sampling biases we will deploy the same number of floats as in (1) but distributed<br />

horizontally and vertically uniformly over the whole Southern Ocean. One of the<br />

main questions is whether Lagrangian statistics provide sufficient information to quantify<br />

possible mixing barriers at fronts, and whether Lagrangian statistics will be similar<br />

in this respect with uniform deployment as compared to when patches of floats are deployed<br />

in between the fronts as in (1) and (2). We will also address the question whether<br />

analysis in isopycnal space leads to similar conclusions compared to a situation in which<br />

the spreading of the floats is quantified with respect to depth interval.<br />

4.1.3 Analysis<br />

The goal of the proposed work is to infer robust Lagrangian diffusivities and spatial variation.<br />

We will use analysis scripts developed by Griesel et al. (009b) who have undertaken<br />

detailed sensitivity analysis. We will address the following questions:<br />

D–9


What are the characteristics of the Lagrangian trajectories in the Southern Ocean of<br />

POP relative to the mean streamlines? Bower and Rossby (1989) found that floats left<br />

the Gulf Stream much earlier at a depth of about 1000 m than closer to the surface, which<br />

could be an indication of enhanced eddy mixing at depth in the core of the Gulf Stream<br />

(see also Bower et al., 1985). However, deep floats were associated with little change<br />

in potential vorticity (PV) along their tracks, and therefore deep cross-stream exchange<br />

was not associated with cross-PV exchange (Bower, 1993). Several studies have examined<br />

mechanisms to explain how cross-stream exchange of fluid parcels is related to the<br />

occurence of meanders in jets (e.g. Bower, 1991; Dutkiewicz et al., 1993; Dutkiewicz and<br />

Paldor, 1994). Since we will archive the four-dimensional Eulerian POP fields, we will<br />

evaluate how cross-frontal exchanges are related to the meander time and space scale, to<br />

the definition of mean streamlines, and to the PV gradients. We will examine the individual<br />

pathways of the floats, and assess how quickly and where floats escape from the ACC<br />

and cross the Subantarctic Front. We will evaluate the length of time that floats remain<br />

within certain dynamical regimes, characterized for example by pressure or speed. These<br />

analyses will provide insights into where property exchanges occur and provide a basis<br />

for the following Lagrangian diffusivity and dispersion evaluation.<br />

Why and when is the diffusive limit reached? Our previous POP float results in the<br />

ACC have demonstrated the need to subtract spatially varying mean flows and the use<br />

of long enough time lags in the integral of the velocity autocovariance (Griesel et al.,<br />

009b). One hypothesis is that with a large enough number of floats the oscillations in<br />

the diffusivity as a function of time lag (Figure 4) arising from rotational components<br />

will cancel each other, and the correct diffusive limit can be reached. On the other hand<br />

Griesel et al. (009a) have concluded that in the case of the Eulerian eddy heat fluxes, the<br />

rotational component is not reduced even if the fluxes are averaged over scales much<br />

larger than the eddy scales. The analysis of Griesel et al. (009b) also showed that upper<br />

ocean trajectories showed more meandering and looping behavior than trajectories from<br />

deeper isopycnal layers. This might be an indication of floats getting locked in with the<br />

mean flow meanders in the upper ocean leading to reduced eddy length scales. We will<br />

analyze in detail the sensitivities of κ ∞ from equation (4) to the definition of mean streamlines,<br />

the subtraction of different means, the number of floats used, the length of the time<br />

lag, and bin sizes. We will evaluate the number of looping and non-looping trajectories<br />

for different depth levels and within and outside the ACC and whether the diffusivity<br />

converges to the same limit if more and more trajectories are used. We will concentrate<br />

on the symmetric components of the diffusivity tensor from equation (4) that represent<br />

irreversible mixing processes but will also compare them to the antisymmetric parts.<br />

What are the robust horizontal and vertical distributions of the Lagrangian eddy diffusivities?<br />

With the proposed POP float deployments we will be able to sample regions<br />

both north and south of the Subantarctic Front. We will investigate the hypothesis that<br />

eddy mixing is suppressed in the core of the ACC whereas it is not supressed in areas<br />

where the strength of the mean flow is comparable to the phase speed of Rossby waves.<br />

Results will be compared to recent estimates from Marshall et al. (2006); Abernathy et al.<br />

(2008) and Naveira Garabato et al. (2009) who use different measures for the effective dif-<br />

D–10


Figure 5: Eddy temperature fluxes v ′ T ′ (a), mean temperature<br />

)<br />

gradients ∂ y T (b), the horizontal<br />

divergence of the eddy temperature flux ∇ h ·<br />

(u ′ h T′ (c) and the Laplacian of mean temperature<br />

∇ 2 T (d) at a depth of 300 m in 1/10 ◦ POP for a region in the Agulhas Retroflection (see Griesel<br />

et al. (009a)).<br />

fusivity. According to Ferrari and Nikurashin (2009) and Naveira Garabato et al. (2009)<br />

the Taylor diffusivity may not be able to accurately capture the reduced mixing where the<br />

mean flow is strong. Here we will investigate the ability of Lagrangian statistics to detect<br />

mixing barriers. This will be achieved by correlating the strength of the mean flow and<br />

the size of the PV gradients with the Lagrangian diffusivities.<br />

4.2 Objective 2: How well do Lagrangian diffusivities parameterize<br />

Eulerian eddy tracer transports and their spatio-temporal distributions<br />

across the ACC?<br />

4.2.1 Background<br />

Davis (1987, 1991) developed the theoretical underpinning, in which way the single particle<br />

Lagrangian diffusivity as a function of time lag is related to the Eulerian eddy fluxes<br />

and mean tracer gradients (see equations 1-2). The term that actually enters the tracer<br />

balance equation is the divergence of the eddy tracer flux<br />

(<br />

)<br />

∇ · 〈u i ′ (x, t)Θ′ (x, t)〉 = ∇ · −κij ∞ (x)∂ x j<br />

Θ(x, t)<br />

(6)<br />

Challenges arise in relating the eddy tracer fluxes from the left-hand-sides of equations<br />

(2) and (6) with the mean tracer gradients from the right-hand-sides of equations (2) and<br />

(6). Eulerian eddy fluxes (Figure 5a) and the divergence of the eddy fluxes (Figure 5b),<br />

are locally not correlated with their respective parameterizations (Figure 5c,d). Accordingly,<br />

eddy diffusivities as calculated from the ratio of flux and mean tracer gradient or<br />

D–11


FT(curl)/[FT(div)+FT(curl)]<br />

1<br />

0.98<br />

0.96<br />

0.94<br />

5 m<br />

0.92 318m<br />

918m<br />

0.9<br />

0.001 0.01 0.1 1<br />

k [cycles/degree longitude]<br />

x<br />

10<br />

Figure 6: A comparison of rotational and divergent<br />

parts of the eddy heat flux in the ACC<br />

in POP as a function of wavenumber and<br />

for three different depth levels in the ACC.<br />

Shown is the spectrum of the curl of the horizontal<br />

eddy heat transport, normalized by the<br />

sum of the spectra of the curl and the divergence<br />

of the horizontal eddy heat transport<br />

PSD(∂ x v ′ T ′ −∂ y u ′ T ′ )<br />

(PSD(∂ x u ′ T ′ +∂ y v ′ T ′ )+PSD(∂ x v ′ T ′ −∂ y u ′ T ′ )) , where<br />

PSD is the power spectral density.<br />

divergence of flux and divergence of mean tracer gradient, are extremely noisy with extremely<br />

high spatial variability and are not usable in climate models (Rix and Willebrand,<br />

1996; Roberts and Marshall, 2000; Nakamura and Chao, 2000; Eden et al., 2007a,b). One<br />

reason for low correlations of the eddy fluxes with their parameterizations and for unphysically<br />

large positive and negative values of the diffusivities is the presence of strong<br />

rotational components. However, it is clear that eddy fluxes systematically do affect the<br />

larger-scale, time-mean state. Griesel et al. (009a) investigated whether averaging over<br />

scales larger than the eddy scales will make the POP eddy heat fluxes more parameterizable<br />

as downgradient diffusion and concluded that the rotational part of the eddy fluxes<br />

as measured by the curl always dominates over the divergent part as measured by the divergence<br />

(Figure 6), no matter how much one averages. Rotational and divergent fluxes<br />

can be separated using a Helmholtz decomposition (e.g. Lau and Wallace, 1979; Roberts<br />

and Marshall, 2000; Drijfhout and Hazeleger, 2001; Jayne and Marotzke, 2002). However,<br />

the Helmholtz decomposition depends on the choice of boundary conditions and is therefore<br />

not unique (Fox-Kemper et al., 2003). It is also computationally fairly expensive.<br />

Alternatively, a rotational component can be identified by considering the eddy variance<br />

equation (MS81, Medvedev and Greatbatch, 2004; Eden et al., 2007a). The eddy<br />

tracer flux can be decomposed into advective, diffusive and rotational components. Adding<br />

or removing a rotational eddy flux does not change the mean tracer budget that involves<br />

the divergence of the eddy flux. To provide a closure, the rotational flux in Medvedev<br />

and Greatbatch (2004) is attributed to the component of the eddy variance flux that is parallel<br />

to mean tracer surfaces. The diffusive eddy tracer flux in turn is associated with the<br />

eddy variance flux that is normal to the mean tracer surfaces. Then, a diffusion coefficient<br />

can be divised that can serve as a diagnostic to the true irreversible eddy mixing taking<br />

place. For example, Eden (2006) calculated eddy diffusivitities by subtracting the MG<br />

rotational flux from the raw eddy fluxes and yielded a more physical representation of<br />

eddy diffusivities by reducing unphysical negative diffusivities. However, as was shown<br />

by Griesel et al. (009a), the isolation of meaningful rotational components based on the<br />

eddy variance equation is challenging. The reason is that the eddy advective terms in the<br />

eddy variance equation are so dominant that it is difficult to extract the part that balances<br />

the conversion term that is related to the irreversible mixing (Wilson and Williams, 2004;<br />

D–12


κ [m 2 /s]<br />

15000<br />

10000<br />

5000<br />

5 m<br />

318 m<br />

579 m<br />

918 m<br />

κ [m 2 /s]<br />

1500<br />

1000<br />

500<br />

5 m<br />

318 m<br />

579 m<br />

918 m<br />

0<br />

0.001 0.01 0.1 1 10<br />

k [cycles/degree longitude]<br />

x<br />

0<br />

0.001 0.01 0.1 1 10<br />

k [cycles/degree longitude]<br />

x<br />

Figure 7: Eulerian diffusivities as a function of zonal wavenumber (or averaging length scale)<br />

from Griesel et al. (009a) for the core of the ACC. Left: Case FG, which refers to diffusivities as<br />

calculated from the raw fluxes and mean temperature gradients. Right: Case DL, which refers to<br />

diffusivities as calculated from the divergence of the eddy fluxes and the divergence of the mean<br />

gradients. Colors represent different depth levels. Case F quasi diffusivities are much larger and<br />

show more wavenumber dependence than Case DL diffusivities. A general decrease with depth is<br />

apparent in both.<br />

Depth [m]<br />

0<br />

500<br />

1000<br />

1500<br />

2000<br />

2500<br />

κ ⊥<br />

0 1000 2000 3000<br />

Figure 8: Lagrangian cross-stream diffusivities<br />

as a function of depth interval. Lines<br />

are along-streamline averages as a function of<br />

mean depth intervals 5 - 500 m, 500 - 1000 m,<br />

1000 - 1500 m, 1500 - 2500 m over three<br />

streamline intervals: south of Polar Front<br />

(red), around Polar Front (blue), north of Subantarctic<br />

Front (green), from South to North).<br />

Cerovecki et al., 2009; Griesel et al., 009a). These triple correlation terms need to be diagnosed<br />

with sufficient numerical accuracy. This means that having high qualilty data to<br />

ensure conservative discretization of the eddy variance equation by taking into account<br />

triple correlation terms is essential. With POP, we will be able to implement diagnostics<br />

of the eddy variance equation and rotational fluxes in a conservative manner online. We<br />

will assess whether diffusivity estimates are indeed more physically plausible when rotational<br />

fluxes based on the eddy variance equation are subtracted from the raw eddy<br />

fluxes. The ultimate goal is to come up with a set of Eulerian diffusivities, that we will be<br />

able to compare to the Lagrangian diffusivities in a consistent way.<br />

Eulerian eddy diffusivities decrease with depth in Griesel et al. (009a) (Figure 7),<br />

whereas the Lagrangian diffusivities are depth invariant in the core of the ACC (Figure<br />

8). The magnitudes of the Lagrangian diffusivities also differ. Eulerian quasi diffusivities,<br />

calculated from the ratio of eddy tracer fluxes and mean gradients, tend to be an order<br />

of magnitude higher than the Lagrangian ones. Eulerian diffusivities calculated from the<br />

ratio of the divergence of the eddy flux and the divergence of the mean gradients, tend<br />

to be about a factor of two smaller than the Lagrangian ones. We plan to explore this discrepancy<br />

which we suspect might be due to different axes orientation and differences in<br />

the role of the rotational parts. Another way of understanding the relationship between<br />

D–13


Eulerian and Lagrangian diagnostics is to assess whether the Eulerian and Lagrangian<br />

eddy scales are related (Middleton, 1985). This will enable us to infer whether the Lagrangian<br />

sampling is in the fixed-float or frozen-field regime (see Lumpkin et al., 2001).<br />

As a float passes through an eddy field, it mixes the field’s spatial and temporal variability<br />

in its time series of displacement. The resulting relationship between the Lagrangian time<br />

scale T L and the Eulerian time scale T E was quantified by Middleton (1985) for an idealized<br />

eddy field. Corrsin’s conjecture expresses the idea that while floats experience both<br />

temporal and spatial decay of the autocorrelation, the Eulerian diffusivities only observe<br />

the temporal decay, leading to Lagrangian eddy scales that are smaller than the Eulerian.<br />

4.2.2 Analysis<br />

The comparison and synthesis of the Lagrangian and Eulerian analyses will be twofold:<br />

(1) We will assess whether the estimated Lagrangian and Eulerian diffusivities have similar<br />

magnitudes and spatial distributions. This comparison will be carried out using both<br />

integrated and more local eddy tracer transports, and will involve evaluation of the elaborated<br />

flux versus gradient law (equation 1). (2) We will compare the different measures<br />

for eddy scales and assess how eddy scales and eddy diffusivities may be parameterized<br />

by large-scale quantities. We ask the following questions:<br />

How do Eulerian and Lagrangian diffusivities compare? We will first characterize the<br />

Eulerian eddy transports. We will assess the zonally and along-streamline averaged eddy<br />

tracer transports for different tracers (temperature, salinity, PV and a passive tracer), and<br />

their horizontal and vertical distribution. The raw eddy tracer fluxes will be compared<br />

to the divergence of the eddy tracer fluxes. We will then calculate eddy diffusivities from<br />

the ratio of the raw fluxes and mean tracer gradients, both averaged over the Lagrangian<br />

bins, and from the divergence of the eddy fluxes and parameterization. We will also<br />

compare these more local estimates to estimates arising when the eddy fluxes are averaged<br />

along circumpolar streamlines, thereby eliminating the rotational part. We ask how<br />

the depth dependence of these Eulerian diffusivities comes about from the depth dependence<br />

of the eddy fluxes and mean gradients, thereby considering that different tracers<br />

may exhibit different spatial distributions. Medvedev and Greatbatch (2004) and Eden<br />

et al. (2007a) have defined methods to compute the rotational components of fluxes by<br />

making use of a conservative discretization of the eddy variance budget. This requires a<br />

large number of terms, and we will therefore estimate the rotational fluxes as part of the<br />

POP model run rather than in the post analysis. We will then ask whether subtracting<br />

rotational components from the raw eddy fluxes produces eddy diffusivities with physically<br />

meaningful magnitudes and spatial patterns. These Eulerian diffusivities resulting<br />

from the subtraction of rotational parts will be compared to the diffusivities computed<br />

from the raw fluxes and gradients. Finally we will evaluate the extent to which the eddy<br />

flux in eq. (1) depends on the history of the mean field gradient. As noted by Davis (1987),<br />

when κ ∞ τ/L 2 c is not small, where κ ∞ is defined in equation (4, τ is the time lag and L c<br />

is a measure of the characteristic eddy length, there will be large fluctuations in the ratio<br />

of instantaneous flux and mean gradient. What are the consequences for the non-local<br />

definition of the Lagrangian diffusivity, and can it be reconciled with the fact that mean<br />

D–14


velocities and mean tracer gradient seem to be changing over scales that are comparable<br />

to the eddy scales?<br />

How do different definitions of eddy scales and their parameterization with large-scale<br />

quantities compare? We will compare the Lagrangian eddy scales as computed from the<br />

integral of the Lagrangian velocity autocovariance with the ones deduced from Eulerian<br />

time series at fixed points. Furthermore, we will calculate classical mixing lengths as defined<br />

by Armi and Stommel (1983) and used by e.g. Naveira Garabato et al. (2009). Can<br />

the classical mixing length be related to the Eulerian and Lagrangian eddy scales? And<br />

is there a difference in their spatial distributions? Finally, we will explore parameterizations<br />

of the eddy diffusivities themselves with the large-scale thermodynamic quantities<br />

that arise from theories of a baroclinically unstable flow field (Green, 1970; Stammer,<br />

1998). Central to these parameterizations is the measure of the growth rate of barolinic<br />

eddies 1/T bc = f / √ R i , where R i is the Richardson number obtained from the large-scale<br />

stratification and shear. The eddy diffusivity can be expressed in the general form as<br />

κ = αL bc T −1<br />

bc<br />

. We will test how well different formulations with different definitions of<br />

length scales L bc that have been proposed (Green, 1970; Stone, 1972; Held and Larichev,<br />

1996) compare to the Eulerian and Lagrangian eddy diffusivity estimates and their spatial<br />

distributions.<br />

4.3 Broader Impacts<br />

This project is motivated by the need to understand the role of ocean circulation in moderating<br />

climate, and by the need to improve climate models. The project will explore<br />

different ways to parameterize mixing coefficients, which will be testable in predictive<br />

climate models. A detailed assessment of Lagrangian methods and their relation to eddy<br />

fluxes and diffusivity estimates from other methods will benefit efforts to analyze data<br />

from floats deployed all over the ocean. In particular, the methods investigated in this<br />

study will contribute to the interpretation of in situ data returned by the 180 RAFOS<br />

floats that are being deployed as part of the Diapycnal and Isopycnal Mixing Experiment<br />

in the Southern Ocean (DIMES). Passive tracer fields, POP float trajectories and other POP<br />

output will be archived and made available to interested research groups. The proposed<br />

work will contribute to early career development of a postdoctoral researcher.<br />

D–15


References<br />

Abernathy, R., J. Marshall, E. Shuckburgh, and M. Mazloff, 2008: Studies of enhanced<br />

isopycnal mixing at deep critical levels in the Southern Ocean. J. Phys. Oceanogr., to be<br />

submitted.<br />

Armi, L. and H. Stommel, 1983: Four views of a portion of the north atlantic subtropical<br />

gyre. J. Phys. Oceanogr., 13, 828 – 857.<br />

Bauer, S., M. S. Swenson, A. Griffa, A. J. Mariano, and K. Owens, 1998: Eddy-mean flow<br />

decomposition and eddy-diffusivity estimates in the tropical Pacific Ocean. 1. Methodology.<br />

J. Geophys. Res., 103 (C13), 30,855 – 30,871.<br />

Berloff, P. S. and J. C. McWilliams, 2002a: Material transport in Oceanic gyres. Part I:<br />

Phenomenology. J. Phys. Oceanogr., 32, 764 – 796.<br />

———, 2002b: Material transport in Oceanic gyres. Part II: Hierarchy of stochastic models.<br />

J. Phys. Oceanogr., 32, 797 – 830.<br />

Böning, C. W., A. Dispert, M. Visbeck, S. R. Rintoul, and F. U. Schwarzkopf, 2008: The<br />

response of the Antarctic Circumpolar Current to recent climate change. Nature Geoscience,<br />

1, 864 – 869, doi: 10.1038/ngeo362.<br />

Bower, A. S., 1991: A simple kinematic mechanism for mixing fluid parcels across a meandering<br />

jet. J. Phys. Oceanogr., 21, 173 – 180.<br />

———, 1993: A closer look at particle exchange in the Gulf Stream. J. Phys. Oceanogr., 24,<br />

1399 – 1418.<br />

Bower, A. S. and H. T. Rossby, 1989: Evidence of cross-frontal exchange processes in the<br />

Gulg Stream based on isopycnal RAFOS float data. J. Phys. Oceanogr., 10, 1177 – 1190.<br />

Bower, A. S., H. T. Rossby, and J. L. Lillibridge, 1985: The Gulf Stream - Barrier or Blender?<br />

J. Phys. Oceanogr., 15, 24 – 32.<br />

Bryden, H. L. and S. A. Cunningham, 2003: How wind-forcing and air-sea heat exchange<br />

determine the meridional temperature gradient and stratification for the Antarctic Circumpolar<br />

Current. J. Geophys. Res., 108 (C8), Art. No. 3275.<br />

Byrne, D. A. and J. L. McClean, 2008: Sea level anomaly signals in the Agulhas Current<br />

region. Geophys. Res. Lett., 35 (L13701), doi:10.1029/2008GL034 584.<br />

Cerovecki, I., R. A. Plumb, and W. Heres, 2009: Eddy transport and mixing in a wind and<br />

buoyancy driven jet on a sphere. J. Phys. Oceanogr., accepted.<br />

Danabasoglu, G. and J. Marshall, 2007: Effects of vertical variations of thickness diffusivity<br />

in an ocean general circulation model. Ocean Modelling, 18, 122 – 141.<br />

Davis, R., 1987: Modelling eddy transport of passive tracers. J. Mar. Res., 45, 635 – 666.<br />

E–1


———, 1991: Observing the general circulation with floats. Deep Sea Res., 38 ((Suppl. 1)),<br />

S531 – S571.<br />

———, 2005: Intermediate-depth circulation of the Iindian and South Pacific Oceans measured<br />

by autonomous floats. J. Phys. Oceanogr., 35, 683 – 707.<br />

Döös, K. and D. Webb, 1994: The Deacon cell and other meridional cells of the Southern<br />

Ocean. J. Phys. Oceanogr., 24, 429 – 442.<br />

Drijfhout, S. S. and W. Hazeleger, 2001: Eddy mixing of potential vorticity versus thickness<br />

in an isopycnic ocean model. J. Phys. Oceanogr., 31 (2), 481 – 505.<br />

Dutkiewicz, S., A. Griffa, and D. B. Olson, 1993: Particle diffusion in a meandering jet.<br />

J. Geophys. Res., 98, 16,487 – 16,500.<br />

Dutkiewicz, S. and N. Paldor, 1994: A note on the mixing enhancement in a meandering<br />

jet due to the interaction with an eddy. J. Phys. Oceanogr., 24, 2,418 – 2,423.<br />

Eden, C., 2006: Thickness diffusivity in the Southern Ocean. Geophys. Res. Lett., 33,<br />

doi:10.1029/2006GL026 157.<br />

Eden, C., R. J. Greatbatch, and D. Olbers, 2007a: Interpreting eddy fluxes.<br />

J. Phys. Oceanogr., 37 (5), 1282 – 1296.<br />

Eden, C., R. J. Greatbatch, and J. Willebrand, 2007b: A diagnosis of thickness fluxes in an<br />

eddy-resolving model. J. Phys. Oceanogr., 37 (3), 727 – 742.<br />

Ferrari, R. and M. Nikurashin, 2009: Suppression of eddy mixing across jets in the southern<br />

ocean. J. Phys. Oceanogr., submitted.<br />

Ferreira, D., J. Marshall, and P. Heimbach, 2005: Estimating eddy stresses by fitting dynamics<br />

to observations using a residual-mean ocean circulation model and its adjoint.<br />

J. Phys. Oceanogr., 35, 1891 – 1910.<br />

Fox-Kemper, B., R. Ferrari, and J. Pedlosky, 2003: On the indeterminacy of rotational and<br />

divergent eddy fluxes. J. Phys. Oceanogr., 33, 478 – 483.<br />

Gent, P. R. and J. C. McWilliams, 1990: Isopycnal mixing in ocean circulation models.<br />

J. Phys. Oceanogr., 20, 150 – 155.<br />

Gille, S. T., 2002: Warming of the southern ocean since the 1950s,. Science, 295, 1275 – 1277.<br />

———, 2008: Decadal-scale temperature trends in the southern hemisphere ocean.<br />

J. Clim., 21 (18), 4749 – 4765.<br />

Gille, S. T., A. Lombrozo, J. Sprintall, G. Stephenson, and R. Scarlet, 2009: Anomalous<br />

spiking in spectra of XCTD temperature profiles. J. Atmos. Ocean. Tech., 26 (6), 1157–<br />

1164.<br />

E–2


Gille, S. T., K. Speer, J. R. Ledwell, and A. C. Naveira Garabato, 2007: Mixing and stirring<br />

in the Southern Ocean. EOS, 88 (39), 382 – 383.<br />

Green, J. S., 1970: Transfer properties of the large-scale eddies and the general circulation<br />

of the atmosphere. Quart. J. Roy. Meteor. Soc., 96, 157 – 185.<br />

Griesel, A., S. T. Gille, J. Sprintall, J. L. McClean, J. H. LaCasce, and M. E. Maltrud, 2009b:<br />

Isopycnal diffusivities in the Antarctic Circumpolar Current inferred from lagrangian<br />

floats in an eddying model. to be submitted to JGR.<br />

Griesel, A., S. T. Gille, J. Sprintall, J. L. McClean, and M. E. Maltrud, 2009a: Assessing eddy<br />

heat flux and its parameterization: A wavenumber perspective from a 1/10 ◦ ocean<br />

simulation. Ocean Modelling, doi:10.1016/j.ocemod.2009.05.004.<br />

Griesel, A. and M. A. Morales Maqueda, 2006: The relation of meridional pressure gradients<br />

to North Atlantic Deep Water volume transport in an Ocean General Circulation<br />

Model. Clim. Dyn., 26 (7-8), 781 – 799.<br />

Griffies, S. M., 1998: The Gent-McWilliams skew flux. J. Phys. Oceanogr., 28, 831 – 841.<br />

Hallberg, R. W. and A. Gnanadesikan, 2006: The role of eddies in determining the structure<br />

and response of the wind-driven Southern hemisphere overturning: Results from<br />

the Modelling Eddies in the Southern Ocean (MESO) project. J. Phys. Oceanogr., 36, 2232<br />

– 2252.<br />

Held, K. M. and V. D. Larichev, 1996: A scaling theory for horizontally homogeneous,<br />

baroclinically unstable flow on a beta plane. J. Atmos. Sci., 53, 946 – 952.<br />

Hogg, A. M., J. P. Meredith, J. R. Blundell, and C. Wilson, 2008: Eddy heat flux in the<br />

Southern Ocean: Response to variable wind forcing. J. Clim., 21 (4), 608 – 620.<br />

Jayne, S. R. and J. Marotzke, 2002: The oceanic eddy heat transport. J. Phys. Oceanogr., 32,<br />

3328 – 3345.<br />

Johnson, G. C. and H. L. Bryden, 1989: On the size of the Antarctic Circumpolar Current.<br />

Deep Sea Res., 36, 39 – 53.<br />

Kamenkovich, I. and E. Sarachik, 2004: Mechanisms controlling the sensitivity of the<br />

Atlantic thermohaline circulation to the parameterization of eddy transports in ocean<br />

GCMs. J. Phys. Oceanogr., 34, 1628.<br />

Karsten, R., H. Jones, and J. Marshall, 2002: The role of eddy transfer in setting the stratification<br />

and transport of a circumpolar current. J. Phys. Oceanogr., 32 (1), 39 – 54.<br />

Karsten, R. and J. Marshall, 2002: Constructing the residual circulation of the ACC from<br />

observations. J. Phys. Oceanogr., 32, 3315 – 3327.<br />

Keffer, T. and G. Holloway, 1988: Estimating Southern Ocean eddy flux of heat and salt<br />

from satellite altimetry. Nature, 332, 624 – 626.<br />

E–3


Krauss, W. and C. Böning, 1987: Lagrangian properties of eddy fields in the northern<br />

North Atlantic as deduced from satellite-tracked buoys. J. Mar. Res., 45, 259 – 291.<br />

LaCasce, J. H., 2008: Lagrangian statistics from oceanic and atmospheric observations.<br />

Transport and Mixing in Geophysical Flows, Weiss, J. B. and A. Provenzale, Eds., Springer.<br />

Lau, N. C. and J. M. Wallace, 1979: On the distribution of horizontal transports by transient<br />

eddies in the Northern Hemisphere wintertime circulation. J. Atmos. Sci., 36, 1844<br />

– 1861.<br />

Lumpkin, R., A.-M. Treguier, and K. Speer, 2001: Lagrangian eddy scales in the Northern<br />

Atlantic Ocean. J. Phys. Oceanogr., 32, 2425 – 2440.<br />

Maltrud, M. E., F. Bryan, M. Hecht, E. Hunke, D. Ivanova, J. L. McClean, and S. Peacock,<br />

2008: Global Ocean Modelline in the eddying regime using POP. CLIVAR Exchanges,<br />

13 (1), 5 – 8.<br />

Maltrud, M. E. and J. L. McClean, 2005: An eddy resolving global 1/10 o ocean simulation.<br />

Ocean Modelling, 8 (1-2), 31 – 54.<br />

Marshall, J., D. Olbers, H. Ross, and D. Wolf-Gladrow, 1993: Potential vorticity constraints<br />

on the dynamics and hydrography of the Southern Ocean. J. Phys. Oceanogr., 23, 456 –<br />

487.<br />

Marshall, J. and T. Radko, 2003: Residual mean solutions for the Antarctic Circumpolar<br />

Current and its associated overturning circulation. J. Phys. Oceanogr., 33 (11), 2341 –<br />

2354.<br />

Marshall, J., E. Shuckburgh, H. Jones, and C. Hill, 2006: Estimates and implications<br />

of surface eddy diffusivity in the Southern Ocean derived from tracer transport.<br />

J. Phys. Oceanogr., 36, 1806 – 1821.<br />

Marshall, J. and G. Shutts, 1981:<br />

J. Phys. Oceanogr., 11, 1677 – 1680.<br />

A note on rotational and divergent eddy fluxes.<br />

McClean, J. L., D. P. Ivanova, and J. Sprintall, 2005: Remote origins of interannual variability<br />

in the Indonesian Throughflow region from data and a global POP simulation.<br />

J. Geophys. Res., 110 (C10013), doi:10.1029/2004JC002 858.<br />

McClean, J. L., S. Jayne, M. E. Maltrud, and D. P. Ivanova, 2008: The fidelity of ocean<br />

models with explicit eddies. Eddy-Resolving Ocean Modelling, Hecht, M. and H. Hasumi,<br />

Eds., AGU Monograph Series, in press.<br />

McClean, J. L., M. L. Maltrud, and F. O. Bryan, 2006: Measures of the fidelity of eddying<br />

ocean models. <strong>Oceanography</strong>, 19 (1), 60 – 73.<br />

Medvedev, A. S. and R. J. Greatbatch, 2004: On advection and diffusion in the mesosphere<br />

and lower thermosphere: The role of rotational fluxes. J. Geophys. Res., 109, D07 104,<br />

doi:10.1029/2003JD003 931.<br />

E–4


Middleton, J., 1985: Drifter spectra and diffusivities. J. Mar. Res., 43, 37 – 55.<br />

Nakamura, M. and Y. Chao, 2000: On the eddy isopycnal thickness diffusivity of the<br />

Gent-McWilliams subgrid mixing parameterization. J. Clim., 13, 502 – 510.<br />

Nakamura, N., 1996: Two-dimensional mixing, edge formation, and permeability in area<br />

coordinates. J. Atmos. Sci., 53, 1524 – 1537.<br />

Naveira Garabato, A. C., R. Ferrari, and K. L. Polzin, 2009: Eddy-induced mixing in the<br />

Southern Ocean. submitted to.<br />

Oh, S. I., V. Zhurbas, and W. S. Park, 2000: Estimating horizontal diffusivity in the East<br />

Sea (Sea of Japan) and the northwest Pacific from satellite-tracked drifter data. J. Geophys.<br />

Res., 105, 6483 – 6492.<br />

Olbers, D. and M. Visbeck, 2005: A model of the zonally averaged stratification and overturning<br />

in the Southern Ocean. J. Phys. Oceanogr., 35, 1190 – 1205.<br />

Phillips, H. E. and S. R. Rintoul, 2000: Eddy variability and energetics from direct<br />

current measurements in the Antarctic Circumpolar Current south of Australia.<br />

J. Phys. Oceanogr., 30, 3050 – 3075.<br />

Prasad, T. G. and J. L. McClean, 2004: Mechanisms for anomalous warming in the<br />

western Indian Ocean during Dipole Mode events. J. Geophys. Res., 109 (C02019),<br />

doi:10.1029/2003JC001 872.<br />

Rainville, L., S. R. Jayne, J. L. McClean, and M. E. Maltrud, 2007: Formation of Subtropical<br />

Mode Water in a high-resolution POP simulation of the Kuroshio extension region.<br />

Ocean Modelling, 17, 338 – 356.<br />

Rix, N. H. and J. Willebrand, 1996: Parameterization of mesoscale eddies as inferred from<br />

a high-resolution circulation model. J. Phys. Oceanogr., 26, 2281 – 2285.<br />

Roberts, M. J. and D. P. Marshall, 2000: On the validity of downgradient eddy closures in<br />

ocean models. J. Geophys. Res., 105 (C12), 28,613 – 28,627.<br />

Sallée, J. B., K. Speer, R. Morrow, and R. Lumpkin, 2008: An estimate of Lagrangian eddy<br />

statistics and diffusion in the mixed layer of the Southern Ocean. J. Mar. Res.<br />

Shuckburgh, E., H. Jones, J. Marshall, and C. Hill, 2008a: Quantifying the eddy diffusivity<br />

of the Southern Ocean I: temporal variability. J. Phys. Oceanogr., submitted.<br />

———, 2008b: Quantifying the eddy diffusivity of the Southern Ocean II: spatial variability<br />

and the role of the mean flow. J. Phys. Oceanogr., submitted.<br />

Smith, K. and J. Marshall, 2009: Evidence for deep eddy mixing in the Southern Ocean.<br />

J. Phys. Oceanogr., 39, 50 – 69.<br />

Smith, R. D., M. E. Maltrud, F. O. Bryan, and M. W. Hecht, 2000: Numerical simulation of<br />

the North Atlantic Ocean at 1/10 o . J. Phys. Oceanogr., 20, 1532 – 1561.<br />

E–5


Speer, K., S. R. Rintoul, and B. M. Sloyan, 2000: The diabatic Deacon cell. J. Phys. Oceanogr.,<br />

30, 3212 – 3222.<br />

Stammer, D., 1998: On eddy characteristics, eddy transports and mean flow properties.<br />

J. Phys. Oceanogr., 28, 727 – 739.<br />

Stone, P. H., 1972: A simplified radiative-dynamical model for the static stability of rotating<br />

atmospheres. J. Atmos. Sci., 29, 405 – 418.<br />

Taylor, G. I., 1921: Diffusion by continuous movements. Proc. Roy. Soc. A, 64, 476 – 490.<br />

Tokmakian, R. and J. L. McClean, 2003: How realistic is the high frequency signal of a<br />

0.1 ◦ ocean model? J. Geophys. Res., 108 (C4), 3115, doi:10.1029/2002JC001 446.<br />

Treguier, A. M., 1999: Evaluating eddy mixing coefficients from eddy-resolving ocean<br />

models: A case study. J. Mar. Res., 57, 89 – 108.<br />

Treguier, A. M., M. H. England, S. R. Rintoul, G. Madec, J. L. Sommer, and J.-M. Molines,<br />

2007: Southern Ocean overturning across streamlines in an eddying simulation of the<br />

Antarctic Circumpolar Current. Ocean Sci., 3, 491 – 507.<br />

Veneziani, M., A. Griffa, Z. D. Garraffo, and E. P. Chassignet, 2005: Lagrangian spin parameter<br />

and coherent structures from trajectories released in a high-resolution ocean<br />

model. J. Mar. Res., 63, 753 – 788.<br />

Veneziani, M., A. Griffa, A. M. Reynolds, and A. J. Mariano, 2004: Oceanic turbulence and<br />

stochastic models from subsurface lagrangian data for the north-west atlantic ocean.<br />

J. Phys. Oceanogr., 34, 1884 – 1906.<br />

Waugh, D. W. and E. R. Abraham, 2008: Stirring in the global surface ocean. Geophys.<br />

Res. Lett., 35, L20 605, doi:10.1029/2008GL035 526.<br />

Wilson, C. and R. G. Williams, 2004: Why are eddy fluxes of potential vorticity difficult to<br />

parameterize ? J. Phys. Oceanogr., 34, 142 – 155.<br />

Zhurbas, V. and S. I. Oh, 2003: Drifter-derived maps of lateral diffusivity in the Pacific and<br />

Atlantic Oceans in relation to surface circulation patterns. J. Geophys. Res., 109, C05 015.<br />

E–6

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!