1. Introduction
Fluctuations on scales of roughly 20–300 km, which derive their energy primarily from the baroclinic instability of the large-scale density field, pervade the global ocean and contain a large fraction of the ocean’s energy (Gill et al. 1974). These fluctuations, known as mesoscale eddies, dominate the dispersion of particles and the mixing of tracers on large space and time scales (Lumpkin and Elipot 2010). Understanding the mixing induced by mesoscale eddies is a problem of both fundamental theoretical interest and practical importance.
The practical importance arises in coarse-resolution ocean climate models, which do not resolve mesoscales. Such models simulate the transport of tracers using a Reynolds-averaged formulation of the tracer conservation equation:
Ocean climate models originally specified the components of
In this study, we demonstrate that eddy diffusivities derived from satellite observations are consistent with a simple formula, whose ingredients are the root-mean-square (rms) eddy velocity urms, which is proportional to the eddy kinetic energy (EKE), the eddy size, the eddy propagation speed, and the large-scale mean flow speed. Together with a prognostic equation for EKE (Eden and Greatbatch 2008; Marshall and Adcroft 2010) and a theory for the eddy sizes, these results can then be used to develop a complete eddy closure. This study is focused on relating observable eddy properties to diagnosed diffusivities; a related study (Bates et al. 2013, manuscript submitted to J. Phys. Oceanogr.) describes how to implement these insights in a complete eddy parameterization.
The central tool in our study is the 20-year record of sea surface height (SSH) observed by satellite altimetry. These data provide both time-mean statistical properties of eddies and a continuous record of surface geostrophic velocity fields. Past authors have used statistical information from the altimeter to estimate eddy diffusivity but lacked a “ground truth” based on an empirical observation of the eddy flux against which to compare their predictions (e.g., Holloway 1986; Stammer 1998). More recent studies have used the altimetric velocities to directly simulate the transport of passive tracers, leading to data-derived estimates of

The study region in the east Pacific between 180° and 130°W. (a) A snapshot of the SSH anomaly field. (b) The zonal-mean flow U (black line), the eddy velocity urms (red line), and the observed eddy phase speed cobs from an eddy-tracking algorithm (blue line).
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1

The study region in the east Pacific between 180° and 130°W. (a) A snapshot of the SSH anomaly field. (b) The zonal-mean flow U (black line), the eddy velocity urms (red line), and the observed eddy phase speed cobs from an eddy-tracking algorithm (blue line).
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1
The study region in the east Pacific between 180° and 130°W. (a) A snapshot of the SSH anomaly field. (b) The zonal-mean flow U (black line), the eddy velocity urms (red line), and the observed eddy phase speed cobs from an eddy-tracking algorithm (blue line).
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1
2. Satellite data and diffusivity calculations
Obtaining direct estimates of mesoscale eddy diffusivity from observations is challenging. The most straightforward method is to construct a local diffusivity based on mooring measurements of the eddy flux of heat (Bryden and Heath 1985). This approach is compromised by the shortness of the available time series (Wunsch 1999) and by the presence of “rotational fluxes,” nondivergent components of the eddy flux vector field that can dominate local measurements (Marshall and Shutts 1981). Lagrangian observations, from surface drifters and subsurface floats, can also be used to estimate diffusivity [review by LaCasce (2008) and Rypina et al. (2012)]. This approach has been more successful but is still limited in accuracy and spatial resolution by the number and location of drifter trajectories (Klocker et al. 2012a,b). In recent years, an efficient, accurate method has been developed that uses real surface velocities (as observed by satellite) in combination with simulated passive tracers to infer eddy diffusivities (Marshall et al. 2006; Abernathey et al. 2010; Ferrari and Nikurashin 2010; Klocker et al. 2012b; Abernathey and Marshall 2013). The tracer-based approach offers global coverage for the entire satellite era, which began in 1992.
Satellite observations of sea surface height anomaly provide weekly geostrophic velocities at the sea surface and resolve the largest mesoscale eddies (Chelton et al. 2007). Eddy properties vary widely across the globe. To simplify our problem somewhat, we focus on a sector in the east Pacific between 180° and 130°W that is mostly free of land and in which eddy properties are relatively homogeneous with longitude. This allows us to focus on the variation of eddy properties and mixing rates as a function of latitude only. A snapshot of the sea surface height anomaly in this sector is shown in Fig. 1, revealing the meridional variations in eddy propagation speed and intensity. This sector was also analyzed by Abernathey and Marshall (2013), and our study builds on that work.
To calculate meridional eddy diffusivities, we simulate the advection of a passive tracer using satellite-derived velocity fields. The velocity dataset we employ is the surface geostrophic velocity anomaly from Archiving, Validation, and Interpretation of Satellite Oceanographic data (AVISO). We use 17 years worth of observations, beginning with 6 January 1993. In addition to the standard geostrophic balance, this dataset employs the empirically validated “equatorial–geostrophic” approximation of Lagerloef et al. (1999) to calculate velocities near the equator (between 5°N and 5°S). The AVISO velocity anomalies are interpolated to a 1/10° grid, and a small correction is applied to remove divergence and to enforce periodicity in longitude. For further details of the data and processing, the reader is referred to Abernathey and Marshall (2013). In addition to the AVISO velocity anomalies, different zonal-mean flows are superimposed, as described in the following section.




Snapshots of passive tracer concentration C after 3 months of advection by AVISO velocities, in three different regions. Note the different color scales of the tracer in each region.
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1

Snapshots of passive tracer concentration C after 3 months of advection by AVISO velocities, in three different regions. Note the different color scales of the tracer in each region.
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1
Snapshots of passive tracer concentration C after 3 months of advection by AVISO velocities, in three different regions. Note the different color scales of the tracer in each region.
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1
3. Mixing length suppression
Mixing length theory



Recent work based on both observations and theoretical arguments has shown that the presence of eddy propagation relative to a background mean flow suppresses eddy diffusivities in the across-current direction (Marshall et al. 2006; Ferrari and Nikurashin 2010; Klocker et al. 2012a,b; Abernathey and Marshall 2013; Tulloch et al. 2013, manuscript submitted to J. Phys. Oceanogr.). This phenomenon can be understood heuristically as follows: if there is no mean flow, and if the eddies are stationary, the eddy stirring acts coherently on the same water masses for a long time, efficiently mixing their properties. But if an individual eddy moves relative to the underlying water, then it no longer acts on the same water masses, and consequently mixing is less efficient. Because urms is independent of mean flow or eddy propagation, this suppression effect must be interpreted as the result of a reduced mixing length.




Summary of the variables used in the manuscript.


This new understanding of mixing length suppression represents a significant advancement in the theory of mesoscale mixing, but so far the issue has been examined only in the Southern Ocean. Furthermore, Eqs. (2) and (3) are only useful for coarse-resolution climate models if the parameters urms, L, cw, and γ can be specified easily based on large-scale fields. Moreover, many simplifying assumptions underlie the theories explained above, and their validity must be tested experimentally. In this study, we test these ideas and their potential limitations in our Pacific sector (Fig. 1) by systematically varying an imposed zonal-mean flow and examining the resulting dependence of Kobs at each latitude. This is a purely kinematic exercise, because the flow fields are not dynamically consistent in general (Haynes et al. 2007).
We run a series of simulations where the tracer C is advected by the velocity field u = u′ + U0i, where u′ is the time-dependent, AVISO-derived eddy velocity; U0 is a constant zonal-mean velocity independent of latitude; and i is the zonal unit vector. The set of 60 simulations employs values of U0 between −0.2 and 0.5 m s−1. Meridional diffusivity Kobs [Eq. (1)] is calculated for each U0 experiment as a function of latitude. The resulting diffusivities are shown as gray shading in Fig. 3a, with examples for particular latitudes plotted in Fig. 3b. First, this shows unequivocally that the diffusivity is highly dependent on the mean flow. Further, as expected from Eqs. (2) and (3), for some value of mean flow (different at each latitude), Kobs reaches a maximum. We identify this maximum diffusivity with the unsuppressed eddy diffusivity

The dependence of eddy diffusivities Kobs on the mean flow. (a) Eddy diffusivities for the east Pacific sector are shown for mean flows ranging from −0.2 to 0.1 m s−1 as gray shading. Overlaid are eddy phase speeds from linear Rossby wave theory including the Doppler shift by the mean flow (red line) and the eddy phase speed estimated in this study (blue line). (b) Examples of the change of Kobs with different mean flows for lat 5°, 10°, 20°, 30°, 40°, and 50°N are presented for diffusivity estimates (solid lines) and the fit of the mixing length formula (dashed lines).
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1

The dependence of eddy diffusivities Kobs on the mean flow. (a) Eddy diffusivities for the east Pacific sector are shown for mean flows ranging from −0.2 to 0.1 m s−1 as gray shading. Overlaid are eddy phase speeds from linear Rossby wave theory including the Doppler shift by the mean flow (red line) and the eddy phase speed estimated in this study (blue line). (b) Examples of the change of Kobs with different mean flows for lat 5°, 10°, 20°, 30°, 40°, and 50°N are presented for diffusivity estimates (solid lines) and the fit of the mixing length formula (dashed lines).
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1
The dependence of eddy diffusivities Kobs on the mean flow. (a) Eddy diffusivities for the east Pacific sector are shown for mean flows ranging from −0.2 to 0.1 m s−1 as gray shading. Overlaid are eddy phase speeds from linear Rossby wave theory including the Doppler shift by the mean flow (red line) and the eddy phase speed estimated in this study (blue line). (b) Examples of the change of Kobs with different mean flows for lat 5°, 10°, 20°, 30°, 40°, and 50°N are presented for diffusivity estimates (solid lines) and the fit of the mixing length formula (dashed lines).
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1
4. Linear Rossby waves versus nonlinear mesoscale eddies
Before interpreting our results, it is helpful to distinguish Rossby waves from nonlinear mesoscale eddies. Linear Rossby waves are zonally propagating oscillations that arise due to the presence of a planetary vorticity gradient. Mesoscale eddies are strongly nonlinear turbulent fluctuations that derive their energy primarily from baroclinic instability. In relation to tracer transport, the most important difference between Rossby waves and nonlinear mesoscale eddies is that eddies can cause irreversible mixing of tracers, whereas linear Rossby waves cannot; distinguishing regions dominated by linear Rossby waves from those of nonlinear mesoscale eddies (or geostrophic turbulence) is therefore important for the interpretation of the results presented in this paper. The analytical model for the eddy diffusivity discussed above [e.g., Eqs. (2) and (3)] applies only to nonlinear eddies. In principle waves and nonlinear eddies of different scales can coexist, given the broad spectrum of variability in the ocean (Wunsch 2010). However, it is evident that at low latitudes, eddy activity is suppressed and energy is transferred instead into waves, while at mid- and high latitudes, eddies are much more prevalent.
Several different approaches have recently been employed to distinguish regions dominated by linear Rossby waves from more nonlinear regions. One such approach is to use the nonlinearity parameter urms/c (Chelton et al. 2007, 2011; Early et al. 2011). (Here c, the intrinsic phase speed, differs from cw, the absolute phase speed; c does not include the Doppler shift by the depth-averaged mean flow.) If urms/c > 1, that is, if the rotational eddy velocity urms exceeds its translational speed c, transforming the coordinates into a co-moving frame will lead to closed streamlines within the eddy. This leads to an eddy with an inner core, which traps fluid and advects it along its path and an outer ring, which continuously entrains and sheds fluid (Early et al. 2011). This outer ring mixes fluid over the eddy size L with the rotational velocity urms, leading to the mixing length arguments in Eq. (2). If urms/c < 1, contours are not closed and the anomalies are more wavelike. Theiss (2004) instead framed his analysis in terms of the Rhines scale LRh and the deformation radius LD; he argued that when LRh/LD > 1, waves are not able to transfer energy into alternating zonal flows, permitting isotropic eddies to exist. This argument implies a critical latitude at LRh/LD = 1. A similar transition from waves to turbulence was demonstrated by Tulloch et al. (2009), who determined the wavelength required to optimally fit the phase speed predicted by linear theory to that observed by altimetry. They concluded that at low latitude there is an overlap between geostrophic turbulence and Rossby wave time scales due to an upscale energy transfer that produces waves, whereas at high latitudes there is no such overlap; that is, waves were not produced and eddies can exist.




The ratio r for the Pacific patch used in this study is shown in Fig. 4a, demonstrating a transition from waves (r < 1) to eddies (r > 1) equatorward of the critical latitude at approximately 18°. Figure 4b shows the number of eddies found per square degree latitude by an eddy-tracking algorithm (Chelton et al. 2011), where an eddy is defined by a closed sea surface height contour. Equatorward of the critical latitude there are almost no eddies observed; presumably instabilities at these low latitudes transfer their energy into alternating zonal flows or waves, rather than producing closed eddies. Figure 4c shows the deformation scale LD and the Rhines scale LRh, which are equal at the critical latitude. One consequence of this transition from waves to turbulence is that we expect different mixing behavior in the two regions. In particular, we do not expect the model of Ferrari and Nikurashin (2010) [Eq. (3)] to be valid in the more linear equatorial region.

Linear Rossby waves vs nonlinear mesoscale eddies and their effect on eddy length scales. (a) The ratio r is shown, with regions r < 1 being dominated by linear Rossby waves, whereas regions with r > 1 are dominated by nonlinear mesoscale eddies. The dashed lines in all panels represent the critical lat at r = 1. (b) Observed eddy numbers per square degree. (c) The empirically calculated eddy size (L; red line), the Rossby radius of deformation (LD; black line), the Rhines scale (LRh; blue line), and the radius of the observed eddy size (Lobs; green line).
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1

Linear Rossby waves vs nonlinear mesoscale eddies and their effect on eddy length scales. (a) The ratio r is shown, with regions r < 1 being dominated by linear Rossby waves, whereas regions with r > 1 are dominated by nonlinear mesoscale eddies. The dashed lines in all panels represent the critical lat at r = 1. (b) Observed eddy numbers per square degree. (c) The empirically calculated eddy size (L; red line), the Rossby radius of deformation (LD; black line), the Rhines scale (LRh; blue line), and the radius of the observed eddy size (Lobs; green line).
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1
Linear Rossby waves vs nonlinear mesoscale eddies and their effect on eddy length scales. (a) The ratio r is shown, with regions r < 1 being dominated by linear Rossby waves, whereas regions with r > 1 are dominated by nonlinear mesoscale eddies. The dashed lines in all panels represent the critical lat at r = 1. (b) Observed eddy numbers per square degree. (c) The empirically calculated eddy size (L; red line), the Rossby radius of deformation (LD; black line), the Rhines scale (LRh; blue line), and the radius of the observed eddy size (Lobs; green line).
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1
5. Eddy phase speeds
Past studies have analyzed the propagation of sea surface height anomalies in the altimetric record (Chelton and Schlax 1996; Chelton et al. 2007, 2011). In recent years, there has been an effort to reconcile such observational estimates with theoretical predictions, and a debate has arisen over what dynamics are observed in altimetric SSH anomaly signals—linear Rossby waves or nonlinear mesoscale eddies (Chelton and Schlax 1996; Killworth et al. 1997, 2004; Chelton et al. 2007; Tulloch et al. 2009; Chelton et al. 2011; O'Brien et al. 2013)? Our experiments can shed light on this debate by providing an independent measurement of the phase speed through an inversion of Eqs. (2) and (3). In the experiments with a range of constant U0 (Fig. 3), we found a maximum diffusivity




6. Eddy length scales
We now turn to understanding the eddy length scales relevant for mixing. To achieve this without the distorting effect of eddy propagation, we use estimates of the unsuppressed eddy diffusivities
The agreement between the observed eddy size Lobs and the empirically calculated eddy size L (using Γ = 0.35) can be seen even more clearly in Fig. 5. This figure shows how eddy length scales depend on the ratio r. Figure 5a plots the ratio l/LD versus r, where l is either Lobs (blue dots) or L (red dots). Figure 5b plots Lobs and L directly against r. Both Figs. 5a and 5b unequivocally show that L agrees very well with Lobs in regions of r > 1, with more complicated behavior in regions of r < 1. In the past, the relationship between L and observed eddy size was more obscure, because previous studies tried to relate suppressed mixing lengths Lmix to the eddy length scales (e.g., Lumpkin et al. 2002; Eden and Greatbatch 2008; Marshall and Adcroft 2010). Only if the suppression effects can be removed, does the mixing length reduce to the length scale of the largest eddies, as has been assumed many times before (Bretherton 1966; Green 1970; Stone 1972; Held and Larichev 1996). This agreement between the observed eddy size and mixing length means we can accurately estimate the unsuppressed eddy diffusivities using

(a) The ratio between eddy length scales and the deformation radius l/LD relative to the ratio r is plotted, with blue dots showing the length scale l = Lobs and red dots showing the length scale l = L. (b) The length scale Lobs (blue dots) and the length scale L (red dots) relative to the ratio r. Note the difference in behavior for the ratio r < 1 (linear waves) and r > 1 (nonlinear eddies), where for r > 1:L = Lobs.
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1

(a) The ratio between eddy length scales and the deformation radius l/LD relative to the ratio r is plotted, with blue dots showing the length scale l = Lobs and red dots showing the length scale l = L. (b) The length scale Lobs (blue dots) and the length scale L (red dots) relative to the ratio r. Note the difference in behavior for the ratio r < 1 (linear waves) and r > 1 (nonlinear eddies), where for r > 1:L = Lobs.
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1
(a) The ratio between eddy length scales and the deformation radius l/LD relative to the ratio r is plotted, with blue dots showing the length scale l = Lobs and red dots showing the length scale l = L. (b) The length scale Lobs (blue dots) and the length scale L (red dots) relative to the ratio r. Note the difference in behavior for the ratio r < 1 (linear waves) and r > 1 (nonlinear eddies), where for r > 1:L = Lobs.
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1

Unsuppressed and suppressed eddy diffusivities in the east Pacific sector. Unsuppressed eddy diffusivities from tracer calculations
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1

Unsuppressed and suppressed eddy diffusivities in the east Pacific sector. Unsuppressed eddy diffusivities from tracer calculations
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1
Unsuppressed and suppressed eddy diffusivities in the east Pacific sector. Unsuppressed eddy diffusivities from tracer calculations
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1
We have employed a mixing efficiency of Γ = 0.35, but what sets this value? One attempt at explaining this parameter was the work of Taylor (1915), which is also the first publication to use a mixing length approach (to explain vertical eddy transport of heat in the atmosphere) known to the authors. Taylor (1915) wrote the vertical eddy diffusivity as
7. The suppression factor
So far we have focused on K0, the unsuppressed diffusivity, using it to understand the relationship between the eddy kinetic energy, size, and mixing in the absence of any suppression effects. But K0 is a hypothetical construction, not the diffusivity experienced in the real ocean. We will now build on these results to understand eddy diffusivities in the presence of a realistic mean flow. To do so, we empirically calculate the meridional diffusivity using the passive tracer with velocity field u = u′ + U(y), where U(y) is the zonally averaged surface zonal flow [here taken from the ocean state estimate of Forget (2010)]. This diffusivity, which we call Kobs, is shown as a black solid line in Fig. 6; it is clearly less than




From Fig. 6, it is clear that Eq. (7) captures the suppression effect qualitatively. The strong meridional variation in Kobs is captured particularly well, with the peaks and valleys located at the right latitudes. However, it is also clear that significant quantitative disagreement in magnitude exists in certain regions. In general, Eq. (7) underestimates Kobs at most latitudes using γL. There are two possibilities to explain this remaining misfit within the context of Eq. (7): either the mixing efficiency Γ now changes its value with latitude or else our choice of γL was incorrect. It is not possible to distinguish these possibilities through our experiments. Because we have no physical model with which to explain a nonconstant Γ, we prefer to treat it as constant and instead focus on γ.
Rather than using γL in Eq. (7), we can find the value that gives the minimum misfit between Kobs and Eq. (7), keeping the other parameters fixed. This leads to an alternate estimate of γ, which we call γfit. In Fig. 7, we plot these two estimates of γ side by side. (Note that the figure plots γ−1, in order to give units of time.) While γL varies between (4 days)−1 and (8 days)−1, γfit is much more constant at approximately (4 days)−1. From this point of view, the fact that γL is in general smaller than γfit is responsible for the overestimated suppression effect in Fig. 6.

Two different estimates of γ. The term γL is the Lagrangian time scale defined in Eq. (8). The term γfit is the value that produces the best agreement between Kobs and Eq. (7).
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1

Two different estimates of γ. The term γL is the Lagrangian time scale defined in Eq. (8). The term γfit is the value that produces the best agreement between Kobs and Eq. (7).
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1
Two different estimates of γ. The term γL is the Lagrangian time scale defined in Eq. (8). The term γfit is the value that produces the best agreement between Kobs and Eq. (7).
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1
The mismatch between γL and γfit suggests a potential shortcoming in the stochastic models of Ferrari and Nikurashin (2010) and Klocker et al. (2012b); these models predict the two quantities should be the same. It is noteworthy that the region of best agreement is the ACC region, where the aforementioned studies were focused. This is one possible reason why the issue was not noticed in those earlier studies. It is well known that the nature of the relationship between Eulerian and Lagrangian scales varies considerably throughout the global ocean (Lumpkin et al. 2002). By examining a wide range of latitudes, our results suggest that some aspects of the model could be refined. Keeping in mind that γ was introduced through a linearization of turbulent nonlinear wave–wave interactions it seems unsurprising that this parameter is a source of uncertainty.
Despite the overestimation of the mixing suppression, we emphasize that, given the wide range of latitudes and Kobs values over this sector, it is evident that Eq. (7) is skillful at reproducing the observed diffusivities and can therefore provide a robust basis for eddy parameterizations. Even employing γL in Eq. (7), the agreement with Kobs is decent. Even better agreement could be achieved by simply using a constant γ = (4 days)−1, although this option is somewhat less satisfying without a deeper explanation for how its value arises.
8. Global extrapolation



Global estimates of eddy diffusivities. Estimates of (a) the unsuppressed eddy diffusivities K0 are compared to (b) the suppressed eddy diffusivities Kmin. Diffusivities are shown on a log10 scale.
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1

Global estimates of eddy diffusivities. Estimates of (a) the unsuppressed eddy diffusivities K0 are compared to (b) the suppressed eddy diffusivities Kmin. Diffusivities are shown on a log10 scale.
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1
Global estimates of eddy diffusivities. Estimates of (a) the unsuppressed eddy diffusivities K0 are compared to (b) the suppressed eddy diffusivities Kmin. Diffusivities are shown on a log10 scale.
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1
One aspect of these global diffusivity estimates merits closer consideration. Recent work on eddy mixing in the ACC has highlighted the importance of “hotspots” for cross-frontal exchange in the lee of topography (Naveira Garabato et al. 2011; Thompson and Sallée 2012). Naveira Garabato et al. (2011) hypothesized that these hotspots are due to the mixing length arguments in Eq. (3) breaking down in these regions because these topographic features lead to a nonzonal shear flow, hence violating the assumptions behind the mixing length model in Eq. (3). To understand these hotspots in the ACC in more detail we show Kmin, EKE, and the inverse of the suppression factor

(a) The suppressed eddy diffusivity log10(Kmin) (m2 s−1), (b) the EKE (m2 s−2), and (c) the inverse of the suppression factor
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1

(a) The suppressed eddy diffusivity log10(Kmin) (m2 s−1), (b) the EKE (m2 s−2), and (c) the inverse of the suppression factor
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1
(a) The suppressed eddy diffusivity log10(Kmin) (m2 s−1), (b) the EKE (m2 s−2), and (c) the inverse of the suppression factor
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1
9. Conclusions
Our goal in this study has been to attempt to explain observed patterns of global surface eddy diffusivity using the mixing length approach. In this case, our “observed” diffusivities were generated from kinematic tracer advection experiments using observed surface geostrophic velocities. Provided one properly accounts for the suppression effect of wave propagation relative to the mean flow, we have demonstrated that knowledge of eddy size, EKE, depth-averaged mean flow, and the Rossby deformation radius (needed to calculate the long-wave Rossby wave phase speed) permits the accurate estimation of spatially varying eddy diffusivities using a simple formula. The resulting eddy diffusivities vary by more than an order of magnitude. This exercise can been seen as an updated version of the studies by Holloway (1986) and Stammer (1998), who also attempted to make estimates of eddy diffusivity from satellite observations. These earlier studies lacked two key elements that we have included here: an updated mixing length theory accounting for the suppression effect and a kinematic tracer simulation to provide a ground truth against which to test the theory’s skill.
In the course of making our estimate and validating the mixing length formulas, we gained two additional insights along the way: 1) nonlinear eddies responsible for mixing travel at the Doppler-shifted long-wave Rossby wave phase speed, and 2) there is a direct relationship between observed eddy size and mixing length, but only for the hypothetical unsuppressed diffusivity. As Eq. (3) makes clear, eddy size is crucial for setting the actual mixing length, but the suppression factor also plays an important role globally (and not just in the Southern Ocean). Another conclusion we reached [confirming work by Theiss (2004) and Tulloch et al. (2009)] is that there exists a critical latitude of approximately ±18° that divides an equatorial regime, dominated by wavelike behavior, from the rest of the ocean, where nonlinear eddies prevail. The mixing length theory we employed applies well to mixing in the nonlinear eddy regime, but not the low-latitude regime. Future work will have to develop a new approach to understand what determines eddy diffusivity in this low-latitude region.
It is important to understand what our study does not do: it does not make a complete closure for eddy diffusivity in terms of the mean state, in the vein of Green (1970), Stone (1972), or Held and Larichev (1996). To make such a closure using our approach, one must go farther and predict the EKE and eddy size based on the mean state. Stammer (1998) illustrates of how this might be done using linear baroclinic instability theory. One limitation of linear theory is its inability to account for the inverse cascade, which causes eddies to grow larger than the deformation radius (Smith 2007). As Fig. 5a illustrates, the strength of the inverse cascade depends on the nonlinearity parameter r. Another limitation of linear stability analysis that treats each profile independently (as in Stammer 1998; Smith 2007; Smith and Marshall 2009; Vollmer and Eden 2013) is that it cannot allow for nonlocal generation and dissipation of EKE. Grooms et al. (2013) recently showed that the eddy energy cycle can in fact be extremely nonlocal, complicating such closure attempts.
Two final ingredients necessary to turn these ideas into a full-fledged closure scheme in a coarse-resolution ocean model are 1) a theory for the vertical structure of the diffusivity and 2) a treatment of the advective part of eddy transport [related to the Gent and McWilliams (1990) parameterization]. Ferrari and Nikurashin (2010) and Klocker et al. (2012a) showed that it is relatively straightforward to extend the mixing length approach in the vertical, provided that the vertical structure of EKE is known. In the interior of the ocean, the mixing length approach would provide an isopycnal diffusivity. Complicating matters, Smith and Marshall (2009) and Abernathey and Marshall (2013) have demonstrated that isopycnal diffusivities have a nontrivial relationship with the Gent–McWilliams coefficient, especially where there is strong vertical structure in these coefficients. A related study (Bates et al. 2013, manuscript submitted to J. Phys. Oceanogr.) addresses this issue through bypassing the Gent–McWilliams approach completely and calculating the eddy-induced velocities in terms of potential vorticity mixing (Marshall 1981).
It is our hope that the ideas presented in this paper can contribute to parameterization efforts by providing simple relations between diffusivity and eddy properties. Beyond this practical goal, we consider these results an incremental step toward better understanding tracer transport by turbulent flows, a problem of fundamental theoretical importance in the ocean and atmosphere. While our results, together with recent studies by Ferrari and Nikurashin (2010) and Klocker et al. (2012a,b), suggest significant progress for extratropical oceanic flows with high zonal symmetry, much remains to be done. The mixing length ideas presented herein clearly fail in the equatorial regime, characterized by values of r < 1. Further research is required to understand how to better describe this regime. The issue is not simply of theoretical interest; the meridional ocean eddy heat transport is quite significant at low latitudes (Jayne and Marotzke 2002). Furthermore, although we have gone to great lengths to empirically validate the mixing length formulas in the Pacific sector, we have not done such a detailed comparison for the full ocean surface flow. Instead, we have simply extrapolated the results from the Pacific sector to a global scale. Although the resulting map is generally consistent with observational estimates (e.g., Rypina et al. 2012; Abernathey and Marshall 2013), understanding the detailed structure of the diffusivity tensor in regions of strong anisotropy, such as western boundary currents, should be a high priority for future research. Additionally, more research needs to be done to understand the variations of mixing efficiency, a topic that has seen little attention so far. Finally, because we have focused on cross-stream diffusivity, we have not attempted to include the effects of shear dispersion in our estimates; this process is expected to enhance diffusivity in the along-jet direction in the presence of mean flow shear (Young and Jones 1991).
Acknowledgments
The authors thank M. Bates, S. Keating, D. Munday, S. Groeskamp, R. Ferrari, M. Nikurashin, M. Spydell, and two anonymous reviewers for very useful comments on the manuscript.
APPENDIX A
Lagrangian Perspective





APPENDIX B
Tracer-Based Mixing Length






We calculated Ktracer from our experiment with (cw − U) = 0, that is, the velocity field that produces unsuppressed diffusivities, to examine the relationship between

Unsuppressed and suppressed eddy diffusivities in the east Pacific sector. Unsuppressed eddy diffusivities from tracer calculations
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1

Unsuppressed and suppressed eddy diffusivities in the east Pacific sector. Unsuppressed eddy diffusivities from tracer calculations
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1
Unsuppressed and suppressed eddy diffusivities in the east Pacific sector. Unsuppressed eddy diffusivities from tracer calculations
Citation: Journal of Physical Oceanography 44, 3; 10.1175/JPO-D-13-0159.1
We now repeat the exercise for suppressed diffusivities using realistic U(y), with both Kobs and Ktracer shown as solid lines in Fig. B1. From this figure it can be seen that the eddy diffusivity estimated from Eq. (B2) is approximately a factor of 2 larger than Kobs. As in Eq. (7), the mixing efficiency itself is quite dependent on the properties of the flow, a conclusion also reached by Thompson and Young (2007) in idealized simulation of baroclinic turbulence. In contrast to the mixing efficiency Γ, we cannot see an obvious relationship of α with the mean flow.
REFERENCES
Abernathey, R., and J. Marshall, 2013: Global surface eddy diffusivities from satellite altimetry. J. Geophys. Res. Oceans, 118, 901–916, doi:10.1002/jgrc.20066.
Abernathey, R., J. Marshall, M. Mazloff, and E. Shuckburgh, 2010: Critical layer enhancement of mesoscale eddy stirring in the Southern Ocean. J. Phys. Oceanogr., 40, 170–184.
Armi, L., and H. Stommel, 1983: Four views of a portion of the North Atlantic subtropical gyre. J. Phys. Oceanogr., 13, 828–857.
Bretherton, F. P., 1966: Critical layer instability in baroclinic flows. Quart. J. Roy. Meteor. Soc., 92, 325–334.
Bryden, H. L., and R. A. Heath, 1985: Energetic eddies at the northern edge of the Antarctic Circumpolar Current. Prog. Oceanogr., 14, 65–87.
Chelton, D. B., and M. G. Schlax, 1996: Global observations of oceanic Rossby waves. Science, 272, 234–238.
Chelton, D. B., M. G. Schlax, R. M. Samelson, and R. A. de Szoeke, 2007: Global observations of large oceanic eddies. Geophys. Res. Lett.,34, L15606, doi:10.1029/2007GL030812.
Chelton, D. B., M. G. Schlax, and R. M. Samelson, 2011: Global observations of nonlinear mesoscale eddies. Prog. Oceanogr., 91, 167–216.
Cushman-Roisin, B., E. P. Chassignet, and B. Tang, 1990: Westward motion of mesoscale eddies. J. Phys. Oceanogr.,20, 758–768.
Danabasoglu, G., and J. Marshall, 2007: Effects of vertical variations of thickness diffusivity in an ocean general circulation model. Ocean Modell., 18, 122–141.
Davis, R., 1987: Modelling eddy transport of passive tracers. J. Mar. Res., 45, 635–666.
Davis, R., 1991: Observing the general circulation with floats. Deep-Sea Res., 38A, S531–S571.
Early, J. J., R. M. Samelson, and D. B. Chelton, 2011: The evolution and propagation of quasigeostrophic ocean eddies. J. Phys. Oceanogr., 41, 1535–1555.
Eden, C., and R. J. Greatbatch, 2008: Towards a mesoscale eddy closure. Ocean Modell., 20, 223–239.
Ferrari, R., and M. Nikurashin, 2010: Suppression of eddy diffusivity across jets in the Southern Ocean. J. Phys. Oceanogr., 40, 1501–1519.
Ferreira, D., J. Marshall, and P. Heimbach, 2005: Estimating eddy stresses by fitting dynamics to observations using a residual-mean ocean circulation model and its adjoint. J. Phys. Oceanogr., 35, 1891–1910.
Forget, G., 2010: Mapping ocean observations in a dynamical framework: A 2004–06 ocean atlas. J. Phys. Oceanogr., 40, 1201–1221.
Fox-Kemper, B., R. Lumpkin, and F. O. Bryan, 2013: Lateral transport in the ocean interior. Ocean Circulation and Climate: A 21st Century Perspective, G. Siedler et al., Eds., Elsevier, 185–209.
Gent, P., and J. McWilliams, 1990: Isopycnal mixing in ocean circulation models. J. Phys. Oceanogr., 20, 150–155.
Gill, A. E., J. S. A. Green, and A. J. Simmons, 1974: Energy partition in the large-scale ocean circulation and the production of mid-ocean eddies. J. Mar. Res., 21, 499–528.
Gille, S. T., and Coauthors, 2012: The Diapycnal and Isopycnal Mixing Experiment: A first assessment. CLIVAR Exchanges, No. 17, International CLIVAR Project Office, Southampton, United Kingdom, 46–48.
Green, J. S. A., 1970: Transfer properties of the large-scale eddies and the general circulation of the atmosphere. Quart. J. Roy. Meteor. Soc., 96, 157–185.
Griffies, S. M., A. Gnanadesikan, R. C. Pacanowski, V. D. Larichev, J. K. Dukowicz, and R. D. Smith, 1998: Isoneutral diffusion in a z-coordinate ocean model. J. Phys. Oceanogr., 28, 805–830.
Grooms, I., L. Nadeau, and S. Smith, 2013: Mesoscale eddy energy locality in an idealized ocean model. J. Phys. Oceanogr., 43, 1911–1923.
Haynes, P., D. A. Poet, and E. Shuckburgh, 2007: Transport and mixing in kinematic and dynamically consistent flows. J. Atmos. Sci., 64, 3640–3652.
Held, I. M., and V. D. Larichev, 1996: A scaling theory for horizontally homogeneous, baroclinically unstable flow on a beta plane. J. Atmos. Sci., 53, 946–953.
Holloway, G., 1986: Estimation of oceanic eddy transports from satellite altimetry. Nature, 323, 243–244.
Jayne, S. R., and J. Marotzke, 2002: The oceanic eddy heat transport. J. Phys. Oceanogr., 32, 3328–3345.
Killworth, P. D., 1986: On the propagation of isolated multilayer and continuously stratified eddies. J. Phys. Oceanogr., 16, 709–716.
Killworth, P. D., D. B. Chelton, and R. A. de Szoeke, 1997: The speed of observed and theoretical long extratropical planetary waves. J. Phys. Oceanogr., 27, 1946–1966.
Killworth, P. D., P. Cipollini, B. M. Uz, and J. R. Blundell, 2004: Physical and biological mechanisms for planetary waves observed in satellite-derived chlorophyll. J. Geophys. Res., 109, C07002, doi:10.1029/2003JC001768.
Klocker, A., R. Ferrari, and J. H. LaCasce, 2012a: Estimating suppression of eddy mixing by mean flows. J. Phys. Oceanogr., 42, 1566–1576.
Klocker, A., R. Ferrari, J. H. LaCasce, and S. T. Merrifield, 2012b: Reconciling float-based and tracer-based estimates of lateral diffusivities. J. Mar. Res., 70, 569–602.
LaCasce, J. H., 2008: Statistics from Lagrangian observations. Prog. Oceanogr., 77, 1–29.
Lagerloef, G., G. Mitchum, R. Lukas, and P. Niiler, 1999: Tropical Pacific near-surface currents estimated from altimeter, wind and drifter data. J. Geophys. Res., 104 (C10), 23 313–23 326.
Lumpkin, R., and S. Elipot, 2010: Surface drifter pair spreading in the North Atlantic. J. Geophys. Res.,115, C12017, doi:10.1029/2010JC006338.
Lumpkin, R., A.-M. Treguier, and K. Speer, 2002: Lagrangian eddy scales in the northern Atlantic Ocean. J. Phys. Oceanogr., 32, 2425–2440.
Marshall, D. P., and A. J. Adcroft, 2010: Parameterization of ocean eddies: Potential vorticity mixing, energetics and Arnold’s first stability theorem. Ocean Modell., 32, 188–204.
Marshall, J., 1981: On the parameterization of geostrophic eddies in the ocean. J. Phys. Oceanogr., 11, 1257–1271.
Marshall, J., and G. J. Shutts, 1981: A note on rotational and divergent eddy fluxes. J. Phys. Oceanogr., 11, 1677–1680.
Marshall, J., A. Adcroft, C. Hill, L. Perelman, and C. Heisey, 1997: A finite-volume, incompressible Navier Stokes model for studies of the ocean on parallel computers. J. Geophys. Res., 102 (C3), 5753–5766.
Marshall, J., E. Shuckburgh, H. Jones, and C. Hill, 2006: Estimates and implications of surface eddy diffusivity in the Southern Ocean derived from tracer transport. J. Phys. Oceanogr., 36, 1806–1821.
McWilliams, J. C., and G. R. Flierl, 1979: On the evolution of isolated, nonlinear vortices. J. Phys. Oceanogr., 9, 1155–1182.
Nakamura, N., 1996: Two-dimensional mixing, edge formation, and permeability diagnosed in an area coordinate. J. Atmos. Sci., 53, 1524–1537.
Naveira Garabato, A. C., R. Ferrari, and K. L. Polzin, 2011: Eddy stirring in the Southern Ocean. J. Geophys. Res.,116, C09019, doi:10.1029/2010JC006818.
O'Brien, R. C., P. Cipollini, and J. R. Blundell, 2013: Manifestation of oceanic Rossby waves in long-term multiparametric satellite datasets. Remote Sens. Environ.,129,111–121.
Prandtl, L., 1925: Bericht über Untersuchungen zur ausgebildeten Turbulenz. Z. Angew. Math. Mech., 5, 136–139.
Redi, M. H., 1982: Oceanic isopycnal mixing by coordinate rotation. J. Phys. Oceanogr., 12, 1154–1158.
Rhines, P. W., 1975: Waves and turbulence on a β-plane. J. Fluid Mech., 69, 417–443.
Rypina, I. I., I. Kamenkovich, P. Berloff, and L. J. Pratt, 2012: Eddy-induced particle dispersion in the near-surface North Atlantic. J. Phys. Oceanogr., 42, 2206–2228.
Sallée, J. B., K. Speer, and R. Morrow, 2008: Southern Ocean fronts and their variability to climate modes. J. Climate, 21, 3020–3039.
Sallée, J. B., K. Speer, and S. R. Rintoul, 2011: Mean-flow and topography control on surface eddy-mixing in the Southern Ocean. J. Mar. Res., 69, 753–777.
Scott, R. B., and F. Wang, 2005: Direct evidence of an oceanic inverse kinetic energy cascade from satellite altimetry. J. Phys. Oceanogr., 35, 1650–1666.
Smith, K. S., 2007: The geography of linear baroclinic instability in Earth’s oceans. J. Mar. Res., 65, 655–683.
Smith, K. S., and J. Marshall, 2009: Evidence for enhanced eddy mixing at middepth in the Southern Ocean. J. Phys. Oceanogr., 39, 50–69.
Stammer, D., 1998: On eddy characteristics, eddy transports, and mean flow properties. J. Phys. Oceanogr., 28, 727–739.
Stone, P. H., 1972: A simplified radiative-dynamical model for the static stability of rotating atmospheres. J. Atmos. Sci., 29, 405–418.
Taylor, G. I., 1915: Eddy motion in the atmosphere. Philos. Trans. Roy. Soc. London, A215, 1–26.
Taylor, G. I., 1921: Diffusion by continuous movements. Proc. London Math. Soc., 20, 196–211.
Theiss, J., 2004: Equatorward energy cascade, critical latitude, and the predominance of cyclonic vortices in geostrophic turbulence. J. Phys. Oceanogr., 34, 1663–1678.
Thompson, A. F., and W. R. Young, 2007: Two-layer baroclinic eddy heat fluxes: Zonal flows and energy balance. J. Atmos. Sci., 64, 3214–3232.
Thompson, A. F., and J.-B. Sallée, 2012: Jets and topography: Jet transitions and the impact on transport in the Antarctic Circumpolar Current. J. Phys. Oceanogr., 42, 956–972.
Tulloch, R., J. Marshall, and K. S. Smith, 2009: Interpretation of the propagation of surface altimetric observations in terms of planetary waves and geostrophic turbulence. J. Geophys. Res.,114, C02005, doi:10.1029/2008JC005055.
Visbeck, M., J. Marshall, and T. Haine, 1997: Specification of eddy transfer coefficients in coarse-resolution ocean circulation models. J. Phys. Oceanogr., 27, 381–403.
Vollmer, L., and C. Eden, 2013: A global map of meso-scale eddy diffusivities based on linear stability analysis. Ocean Modell., 72, 198–209.
Wunsch, C., 1999: Where do ocean eddy heat fluxes matter? J. Geophys. Res., 104 (C6), 13 235–13 249.
Wunsch, C., 2010: Toward a midlatitude ocean frequency–wavenumber spectral density and trend determination. J. Phys. Oceanogr., 40, 2264–2282.
Young, W. R., and S. Jones, 1991: Shear dispersion. Phys. Fluids, 3A, 1087–1101.
Zhang, H.-M., M. D. Prater, and T. Rossby, 2001: Isopycnal Lagrangian statistics from the North Atlantic Current RAFOS float observations. J. Geophys. Res.,106 (C7), 13 817–13 836.
Zhurbas, V., and I. S. Oh, 2003: Lateral diffusivity and Lagrangian scales in the Pacific Ocean as derived from drifter data. J. Geophys. Res., 108, 3141, doi:10.1029/2002JC001596.
Zhurbas, V., and I. S. Oh, 2004: Drifter-derived maps of lateral diffusivity in the Pacific and Atlantic Oceans in relation to surface circulation patterns. J. Geophys. Res.,109, C05015, doi:10.1029/2003JC002241.
By absolute eddy phase speed cw, we mean the phase speed that would be measured by a stationary observer. For baroclinic Rossby waves, this phase speed is determined by a dispersion relation that includes a Doppler shift by the depth-averaged zonal-mean flow
Whereas the depth-averaged zonal-mean flow very accurately explains the Doppler shift of mesoscale eddies, this is not the case for the meridional phase speed of eddies; see Klocker and Marshall (2013, manuscript submitted to Geophys. Res. Lett.) for details. Nevertheless, describing the meridional phase speed with the depth-averaged meridional-mean flow has negligible consequences for the estimates of eddy mixing.