1. Introduction
Oceanic mesoscale eddies play a key role in the global ocean circulation, oceanic ecosystems and the climate system as a whole. Eddies mix, transport, and store tracers such as heat, salt, carbon, oxygen, and nutrients (Lee et al. 2007; Gruber et al. 2011; Gnanadesikan et al. 2013, 2015, 2017; Stewart et al. 2018; Busecke and Abernathey 2019; Jones and Abernathey 2019; Groeskamp et al. 2019). However, mesoscale eddies occur on spatial scales of 10–100 km, which is on the same order or smaller than the horizontal grid resolution of most global climate models (Eden 2007; Chelton et al. 2011; Hallberg 2013; LaCasce and Groeskamp 2020; Martínez-Moreno et al. 2022). Therefore, mesoscale mixing processes are often not explicitly resolved in climate simulations, and instead need to be parameterized (Eden and Greatbatch 2008; Hallberg 2013; Jansen et al. 2015; Zanna et al. 2017; Fox-Kemper et al. 2019; Wang and Stewart 2020). Parameterization of eddy mixing is typically done via an eddy diffusivity
Significantly, climate models are very sensitive to the choice of the diffusivity value (e.g., Ferreira et al. 2005; Pradal and Gnanadesikan 2014; Gnanadesikan et al. 2015; Kjellsson and Zanna 2017; Jones and Abernathey 2019; Holmes et al. 2022; Mak et al. 2022b). In simulations of Earth’s climate, an approximately fivefold increase in the value of
Many parameterizations of
Holloway and Kristmannsson (1984) and Holloway (1986) suggested that if eddies have Rossby wave characteristics, the eddy diffusivity is suppressed. This suppression effect was later shown analytically by e.g., Ferrari and Nikurashin (2010) (using a passive tracer approach), Klocker et al. (2012) (using a Lagrangian approach), and Griesel et al. (2015) (using linear stability analysis). In all of these studies, eddy fields are represented as statistically forced and linearly damped Rossby waves, and it is shown that the cross-stream mixing length is effectively reduced in the presence of a background mean flow if the eddies are propagating relative to the mean flow.
A kinematic interpretation of the suppression mechanism is that the mean flow will advect tracers through the eddy field before the eddy field has had time to mix the tracers in the cross-stream direction. If the eddies did not have an intrinsic phase speed, they would move with the mean flow and thus be able to effectively mix the tracers. The parameterization of Ferrari and Nikurashin (2010) has been widely used in idealized models (Nakamura and Zhu 2010b; Eden 2011; Srinivasan and Young 2014; Kong and Jansen 2017; Wolfram and Ringler 2017; Seland et al. 2020) and validated and applied to the Antarctic Circumpolar Current (Naveira Garabato et al. 2011; Sallée et al. 2011; Meredith et al. 2012; Pennel and Kamenkovich 2014; Chen et al. 2015; Roach et al. 2016; Chapman and Sallée 2017), the Kuroshio Extension (Chen et al. 2014), the Gulf Stream (Bolton et al. 2019), the Nordic seas (Isachsen and Nøst 2012), eastern boundary currents (Bire and Wolfe 2018), and the global ocean (Bates et al. 2014; Klocker and Abernathey 2014; Roach et al. 2018; Busecke and Abernathey 2019; Canuto et al. 2019; Groeskamp et al. 2020).
A different interpretation from the kinematic explanation described above is that the suppression is a dynamical effect caused by gradients in potential vorticity (PV). Marshall et al. (2006) estimated surface eddy diffusivities in the Southern Ocean from satellite altimetry, and found that regions of high and low diffusivity coincide with regions of weak and strong PV gradients, respectively. They suggested that strong PV gradients impose a barrier on lateral transport, inhibiting cross-stream diffusivity. This effect is also observed in the atmosphere (e.g., Dritschel and McIntyre 2008). Nakamura and Zhu (2010b), Klocker et al. (2012), Srinivasan and Young (2014) and Balwada et al. (2016) explicitly linked the mixing barriers caused by PV gradients to the parameterization of Ferrari and Nikurashin (2010) by noting that the PV gradient determines the Rossby wave phase speed; hence, it is the PV gradient that enables the eddies to move relative to the mean flow, which leads to the suppression of the cross-stream eddy diffusivity.
Previous studies have mainly focused on PV gradients caused by the planetary β effect (latitudinal variations of the Coriolis parameter). However, an important factor that should also be considered is the effect of topography. If it is indeed the PV gradient that causes the suppression effect, then we must consider the role of topography as well, because topographic slopes contribute significantly to (barotropic) PV gradients (LaCasce and Speer 1999; LaCasce 2000) and permit topographic Rossby waves (Rhines 1970; Csanady 1976; Hogg 2000). Hence, topographic slopes can also be expected to modulate eddy diffusivity (Jansen et al. 2015). Isachsen (2011) diagnosed eddy diffusivities for different bottom slopes from numerical simulations, and found that the diffusivities were highest for flat bottoms, suggesting a suppression effect of topographic slopes. A relevant question then is how exactly does topography modulate eddy diffusivities, and how to parameterize topographic effects related to eddy mixing. Since topography steers currents, its effects might already be included in the mean flow term from Ferrari and Nikurashin (2010), but only implicitly.
Some recent studies have aimed to express the eddy diffusivity explicitly in terms of topographic slopes using numerical model data. Diagnostic expressions were derived from high-resolution simulations by Brink (2012), Brink and Cherian (2013) and Brink (2016), by Wang and Stewart (2020) and Wei et al. (2022) for buoyancy diffusivity specifically, and by Wei and Wang (2021) for isopycnal diffusivity specifically. Moreover, Nummelin and Isachsen (2024) and Wei et al. (2024) derived parameterizations for the buoyancy diffusivity over topographic slopes and tested them in prognostic coarse-resolution simulations. All of these studies derived parameterizations for the eddy diffusivities using various scaling estimates for the mixing length combined with empirical “suppression” functions. Although the suppression functions from the aforementioned studies perform well in representing suppression of eddy diffusivity by topographic slopes, they are essentially empirical fits to functions that have little dynamical justification. Therefore, the aim of this study is to derive an analytical expression for the suppression of
The rest of this article is organized as follows. In section 2, we derive the analytical model. In section 3, we compare the analytical expression for
2. Theory
Equations (9) and (12) both describe suppression of eddy mixing, but offering different interpretations. Equation (9) expresses suppression in terms of the mean flow and the eddy phase speed, while (12) expresses the suppression in terms of the PV gradient directly. The “velocity formulation” (9) has been used frequently before (Ferrari and Nikurashin 2010; Klocker et al. 2012), while the “PV formulation” (12) was noted by Nakamura and Zhu (2010a) and Klocker et al. (2012). We use the PV form, recognizing that the mean velocity should drop out of the problem, due to the Galilean invariance noted earlier.
3. Validating theory in an idealized channel model
a. Numerical model description
We use the Bergen Layered Ocean Model (BLOM), the ocean component of the Norwegian Earth System Model (NorESM; Seland et al. 2020), in an idealized channel configuration. The simulations are described in Nummelin and Isachsen (2024) and we only give a brief summary here.
The model uses 51 isopycnal levels (potential density referenced to 2000 dbar) with a two-level bulk mixed layer at the surface. The channel configuration is 416 km long in the zonal (x) direction and 1024 km wide in the meridional (y) direction with a 2-km resolution, and is re-entrant in the zonal direction. There are continental slopes of 2000 m extension from the shelf break at 250 m to the bottom of the domain at 2250-m depth on both sides of the channel, centered at 150 km from the domain edge. To trigger instabilities, we add white noise to the bottom topography with a standard deviation of 10 m. The channel is set up on the Northern Hemisphere f plane. The model is initialized from rest with constant salinity and a horizontally homogeneous temperature profile. The density is determined by temperature alone, which has a maximum at the surface and decays exponentially toward the bottom. There is no buoyancy forcing (nor restoring) and we only force the flow with a constant westward wind stress. The surface mixed layer is kept shallow by parameterization of submesoscale mixed layer eddies (Fox-Kemper et al. 2008) that counter the vertical mixing induced by the constant wind forcing.
The wind forcing drives a northward surface Ekman transport. Ekman divergence in the south and convergence in the north drive a westward mean flow U(y). The mean flow is retrograde with respect to topographic waves in the south whereas it is prograde in the north. Upwelling in the south establishes isopycnals that are sloping with the topography whereas downwelling in the north sets up isopycnals that slope against the topography. The tilted isopycnals in both regions are baroclinically unstable, creating an eddy field. Figure 1 shows a snapshot of the fields from one of the simulations.
Illustration of the channel model configuration. The surface elevations represent an exaggerated snapshot of daily SSH anomalies, with the colors showing a snapshot of daily SST anomalies. The purple hues show the zonal mean velocity.
Citation: Journal of Physical Oceanography 54, 5; 10.1175/JPO-D-23-0142.1
We run nine experiments, varying the initial stratification and the width of the continental slope, i.e., the slope angle. The slope aspect ratio α = (slope height)/(slope width) varies between 0.016 and 0.027. These values are fairly representative for continental slopes (LaCasce 2017). All simulations are spun up for 10 years, to a semi-equilibrium where the kinetic energy has stabilized but still has some variability. The model fields are then diagnosed over an additional 5-yr period (between years 11 and 15). The parameter settings and experiments are laid out in Tables 1 and 2, respectively.
BLOM model constants for the channel simulations.
Channel model experiments. The term LRossby is the mean deformation radius [(14)] averaged over the last 5 years of the 15-yr-long experiments in the central basin (where the bottom depth is larger than 2250 m).
b. Computing diffusivities from the model data
The goal is to compare cross-slope eddy diffusivities diagnosed from the model with the parameterizations from (16) and (17). We diagnose diffusivities using the flux-gradient relation
c. Comparing parameterized and diagnosed diffusivities
Figure 2 shows the diagnosed and parameterized cross-slope diffusivities across the channel for all nine experiments (Table 2). The diagnosed diffusivities
Zonal and time mean depth-averaged cross-slope diffusivities across the channel for all nine experiments from Table 2. Diffusivities are all plotted on a logarithmic scale. The continuous black line shows the diagnosed temperature diffusivity, the dashed black line shows the diagnosed PV diffusivity, the purple line shows the parameterized unsuppressed diffusivity from (10), and the red and blue lines show the parameterized diffusivities from (20) and (21), respectively. The parameterized diffusivities are all shown for γ−1 = 4 days. The midbasin part between 400 and 600 km is not shown; here the diffusivity is approximately constant. The gray shaded areas indicate the topographic slopes.
Citation: Journal of Physical Oceanography 54, 5; 10.1175/JPO-D-23-0142.1
Next, the purple, red and blue lines show the parameterized diffusivities
Figure 3 explores the relevance of the anisotropy factor
Diagnosed and parameterized diffusivities (as in Fig. 2) for experiment 5 from Table 2. Diffusivities are all plotted on a logarithmic scale. The dashed lines show the parameterized diffusivities with a constant anisotropy factor
Citation: Journal of Physical Oceanography 54, 5; 10.1175/JPO-D-23-0142.1
Figure 4 shows the profiles of the parameterized diffusivities
Parameterized diffusivities
Citation: Journal of Physical Oceanography 54, 5; 10.1175/JPO-D-23-0142.1
4. Discussion
a. Relevance of eddy length and time scales
In the derivation of (12), it is assumed that mesoscale eddies can be described as monochromatic waves with all energy at a single wavenumber. In reality, the oceanic eddy field contains motions over a broad range of wavenumbers (Wunsch 2010; Wortham and Wunsch 2014). A number of studies have derived eddy diffusivity parameterizations for multichromatic waves. Chen et al. (2015) developed a multiwavenumber theory for eddy diffusivities, but only considered wavenumbers in the along-stream direction. Kong and Jansen (2017) considered the full two-dimensional EKE spectrum to compute eddy diffusivities, but assumed isotropy. Instead, like most other studies, we retained the assumption of monochromatic waves. We considered two different length scales to set the dominant wavelength: the Rossby radius and the topographic Rhines scale. As seen in Fig. 2, we find good agreement between theory and model results in both cases. This suggests that the assumption of monochromatic waves works well with a realistic value for the most energetic eddy length scale for this model.
Other studies have not yet provided a clear conclusion on which length scale best represents eddy mixing length over topographic slopes. Wang and Stewart (2020) and Wei et al. (2022) found that the topographic Rhines scale works well to parameterize eddy diffusivity over retrograde slopes, but that it is not suitable for prograde slopes in a stratified ocean. On the other hand, Wei and Wang (2021) parameterized diffusivities over retrograde slopes using the Rossby radius, and found that the topographic Rhines scale led to an overestimation of the diagnosed diffusivity. These findings suggest differences in eddy length scales between prograde and retrograde slopes. In our simulations, a relevant difference between the two slopes is that EKE is enhanced over the northern (prograde) slope due to Ekman downwelling, but weakened in the south due to Ekman upwelling. Over the southern (retrograde) slope, the suppression of the eddy diffusivity is already captured quite well by
Regarding the eddy velocity decorrelation time scale, there are, to the best of our knowledge, no observational studies on the values of γ in the ocean. As noted, γ is typically left as an adjustable parameter when computing eddy diffusivities (e.g., Klocker and Abernathey 2014; Groeskamp et al. 2020). The value of γ could possibly be inferred from an inverse method, like that employed in Mak et al. (2022a) for the mesoscale eddy energy dissipation time scale. Another option could be to determine γ from the autocorrelation of observational velocity time series. Our results suggest γ varies depending on the relevant dynamics in a region, and this should be examined further.
b. Relation with empirical expressions for eddy diffusivity
c. Challenges for implementation in coarse-resolution climate models
One of the main reasons to study eddy diffusivities over topographic slopes is to create better parameterizations for coarse-resolution climate models. Using (13) with the appropriate length scale to compute eddy diffusivities requires knowledge on the eddy kinetic energy
d. Applicability of results for observations
The skill of parameterizations (20) and (21) in reproducing eddy diffusivities in a numerical model also motivates application to observational data. Direct observations of mesoscale eddy mixing can be made in tracer release experiments (Ledwell et al. 1993, 1998; Tulloch et al. 2014; Zika et al. 2020; Bisits et al. 2023), but these experiments are expensive and labor intensive, and only provide information about a specific region. By contrast, our parameterizations could be used to infer eddy diffusivity values from more easily attainable observations. Groeskamp et al. (2020) applied the velocity formulation from Ferrari and Nikurashin (2010) to an observation-based gridded ocean climatology to create full-depth global estimates of eddy diffusivities. However, the expression of Ferrari and Nikurashin (2010) does not include effects of topographic PV gradients. Moreover, it requires fitting of the eddy decorrelation time scale γ, and approximating the total eddy phase speed cw. Table 4 of Wei and Wang (2021) summarizes the methods that different studies used to determine cw, which include empirical fits to numerical model results (e.g., Klocker et al. 2012; Pennel and Kamenkovich 2014), linear stability analysis (e.g., Eden 2011; Griesel et al. 2015), the use of SSH measurements (e.g., Ferrari and Nikurashin 2010; Naveira Garabato et al. 2011; Sallée et al. 2011; Abernathey and Marshall 2013; Bates et al. 2014; Klocker and Abernathey 2014; Balwada et al. 2016; Roach et al. 2016, 2018; Bolton et al. 2019; Busecke and Abernathey 2019; Groeskamp et al. 2020), or simply assuming cw ≈ 0 (e.g., Meredith et al. 2012; Bire and Wolfe 2018). By contrast, in the PV formulation the term cw − U is replaced by c, the intrinsic eddy phase speed, for which we have an analytical expression in terms of the background (planetary and topographic) PV gradient. Thus, in the barotropic case we can calculate c in a straightforward way from β and the topographic slope, and circumvent the problem of having to determine cw. In the end, the only observational measurements that our (20) and (21) require are information on stratification, topographic slopes, and flow velocity time series (for
5. Summary and conclusions
We derived an analytical expression, (13), to describe depth-averaged eddy diffusivities over topographic slopes (Fig. 2). This expression is a specific case of the general (12) for the cross-stream eddy diffusivity in the presence of a background PV gradient. Equation (12) explicitly links eddy diffusivity to the PV gradient (Nakamura and Zhu 2010b), thus providing a PV formulation of mixing suppression, as opposed to the velocity formulation [(9)] presented in previous studies (e.g., Ferrari and Nikurashin 2010; Klocker et al. 2012). An advantage of the PV formulation is that it does not require information on cw, the Doppler-shifted or apparent phase speed of the eddies, and U, the background mean flow. We circumvent the problem of having to determine cw and U and instead keep an analytical expression for the intrinsic eddy phase speed, which is linked to the PV gradient. Furthermore, keeping the PV gradient ∇Q in the expression for
A number of issues still remain to be addressed. Closures for the eddy anisotropy, EKE, and decorrelation time scale are missing; the physical mechanisms setting the eddy length scale in different dynamical regimes require further study; and the parameterizations for
Acknowledgments.
M. F. S.: Conceptualization, methodology, formal analysis, writing—original draft. J. H. L.: Conceptualization, methodology. S. G.: Conceptualization, methodology, supervision, project administration, funding acquisition. A. N.: Software, formal analysis, resources. P. E. I.: Resources. M. L. J. B.: Supervision, project administration, funding acquisition. Everyone: Writing—review and editing. M. F. S. was funded by the UU-NIOZ Project “The intermittency of large-scale ocean mixing” (Project NZ4543.3). J. H. L. was supported under Project 302743 (the Rough Ocean) of the Norwegian Research Council. A. N. and P. E. I. were funded by the two Research Council of Norway projects KeyClim (295046) and TopArctic (314826). M. L. J. B. was funded by the program of the Netherlands Earth System Science Centre (NESSC), financially supported by the Ministry of Education, Culture and Science (OCW, Grant 024.002.001). The model simulations and storage were performed on resources provided by Sigma2—the National Infrastructure for High Performance Computing and Data Storage in Norway under the Accounts NN9252K, NS9252K, NN9869K, and NS9869K. We thank Julian Mak and one anonymous reviewer for their thorough and helpful reviews of the manuscript.
Data availability statement.
The model configuration and namelists needed for reproducing the results are published in Zenodo (Nummelin 2023b) and available at https://doi.org/10.5281/zenodo.8227381. The key model outputs (Nummelin 2023a) needed for reproducing the analysis are published at the NIRD research data archive and available at https://archive.sigma2.no/pages/public/datasetDetail.jsf?id=10.11582/2023.00129. Scripts for data processing and plotting can be shared upon request.
APPENDIX
Derivation of Expression for Cross-Stream Eddy Diffusivity
REFERENCES
Abernathey, R. P., and J. Marshall, 2013: Global surface eddy diffusivities derived from satellite altimetry. J. Geophys. Res. Oceans, 118, 901–916, https://doi.org/10.1002/jgrc.20066.
Balwada, D., K. G. Speer, J. H. LaCasce, W. B. Owens, J. Marshall, and R. Ferrari, 2016: Circulation and stirring in the southeast Pacific Ocean and the Scotia Sea sectors of the Antarctic Circumpolar Current. J. Phys. Oceanogr., 46, 2005–2027, https://doi.org/10.1175/JPO-D-15-0207.1.
Bates, M., R. Tulloch, J. Marshall, and R. Ferrari, 2014: Rationalizing the spatial distribution of mesoscale eddy diffusivity in terms of mixing length theory. J. Phys. Oceanogr., 44, 1523–1540, https://doi.org/10.1175/JPO-D-13-0130.1.
Bire, S., and C. L. P. Wolfe, 2018: The role of eddies in buoyancy-driven eastern boundary currents. J. Phys. Oceanogr., 48, 2829–2850, https://doi.org/10.1175/JPO-D-18-0040.1.
Bisits, J. I., G. J. Stanley, and J. D. Zika, 2023: Can we accurately quantify a lateral diffusivity from a single tracer release? J. Phys. Oceanogr., 53, 647–659, https://doi.org/10.1175/JPO-D-22-0145.1.
Boland, E. J. D., A. F. Thompson, E. Shuckburgh, and P. H. Haynes, 2012: The formation of nonzonal jets over sloped topography. J. Phys. Oceanogr., 42, 1635–1651, https://doi.org/10.1175/JPO-D-11-0152.1.
Bolton, T., R. Abernathey, and L. Zanna, 2019: Regional and temporal variability of lateral mixing in the North Atlantic. J. Phys. Oceanogr., 49, 2601–2614, https://doi.org/10.1175/JPO-D-19-0042.1.
Brink, K. H., 2012: Baroclinic instability of an idealized tidal mixing front. J. Mar. Res., 70, 661–688, https://doi.org/10.1357/002224012805262716.
Brink, K. H., 2016: Continental shelf baroclinic instability. Part I: Relaxation from upwelling or downwelling. J. Phys. Oceanogr., 46, 551–568, https://doi.org/10.1175/JPO-D-15-0047.1.
Brink, K. H., 2017: Surface cooling, winds, and eddies over the continental shelf. J. Phys. Oceanogr., 47, 879–894, https://doi.org/10.1175/JPO-D-16-0196.1.
Brink, K. H., and D. A. Cherian, 2013: Instability of an idealized tidal mixing front: Symmetric instabilities and frictional effects. J. Mar. Res., 71, 425–450, https://doi.org/10.1357/002224013812587582.
Busecke, J. J. M., and R. P. Abernathey, 2019: Ocean mesoscale mixing linked to climate variability. Sci. Adv., 5, eaav5014, https://doi.org/10.1126/sciadv.aav5014.
Buzzicotti, M., B. A. Storer, H. Khatri, S. M. Griffies, and H. Aluie, 2023: Spatio-temporal coarse-graining decomposition of the global ocean geostrophic kinetic energy. J. Adv. Model. Earth Syst., 15, e2023MS003693, https://doi.org/10.1029/2023MS003693.
Canuto, V. M., Y. Cheng, A. M. Howard, and M. S. Dubovikov, 2019: Three-dimensional, space-dependent mesoscale diffusivity: Derivation and implications. J. Phys. Oceanogr., 49, 1055–1074, https://doi.org/10.1175/JPO-D-18-0123.1.
Chapman, C., and J.-B. Sallée, 2017: Isopycnal mixing suppression by the Antarctic Circumpolar Current and the Southern Ocean meridional overturning circulation. J. Phys. Oceanogr., 47, 2023–2045, https://doi.org/10.1175/JPO-D-16-0263.1.
Charney, J. G., 1947: The dynamics of long waves in a baroclinic westerly current. J. Meteor., 4, 136–162, https://doi.org/10.1175/1520-0469(1947)004<0136:TDOLWI>2.0.CO;2.
Chelton, D. B., R. A. deSzoeke, M. G. Schlax, K. El Naggar, and N. Siwertz, 1998: Geographical variability of the first baroclinic Rossby radius of deformation. J. Phys. Oceanogr., 28, 433–460, https://doi.org/10.1175/1520-0485(1998)028<0433:GVOTFB>2.0.CO;2.
Chelton, D. B., M. G. Schlax, and R. M. Samelson, 2011: Global observations of nonlinear mesoscale eddies. Prog. Oceanogr., 91, 167–216, https://doi.org/10.1016/j.pocean.2011.01.002.
Chen, R., J. L. McClean, S. T. Gille, and A. Griesel, 2014: Isopycnal eddy diffusivities and critical layers in the Kuroshio Extension from an eddying ocean model. J. Phys. Oceanogr., 44, 2191–2211, https://doi.org/10.1175/JPO-D-13-0258.1.
Chen, R., S. T. Gille, J. L. McClean, G. R. Flierl, and A. Griesel, 2015: A multiwavenumber theory for eddy diffusivities and its application to the southeast Pacific (DIMES) region. J. Phys. Oceanogr., 45, 1877–1896, https://doi.org/10.1175/JPO-D-14-0229.1.
Chouksey, A., A. Griesel, M. Chouksey, and C. Eden, 2022: Changes in global ocean circulation due to isopycnal diffusion. J. Phys. Oceanogr., 52, 2219–2235, https://doi.org/10.1175/JPO-D-21-0205.1.
Csanady, G. T., 1976: Topographic waves in Lake Ontario. J. Phys. Oceanogr., 6, 93–103, https://doi.org/10.1175/1520-0485(1976)006<0093:TWILO>2.0.CO;2.
Davis, R. E., 1987: Modeling eddy transport of passive tracers. J. Mar. Res., 45, 635–666, https://doi.org/10.1357/002224087788326803.
Davis, R. E., 1991: Observing the general circulation with floats. Deep-Sea Res., 38A, S531–S571, https://doi.org/10.1016/S0198-0149(12)80023-9.
Delsole, T., 2004: Stochastic models of quasigeostrophic turbulence. Surv. Geophys., 25, 107–149, https://doi.org/10.1023/B:GEOP.0000028160.75549.0d.
Dijkstra, H. A., 2008: Dynamical Oceanography. Springer, 407 pp.
Dritschel, D. G., and M. E. McIntyre, 2008: Multiple jets as PV staircases: The Phillips effect and the resilience of eddy-transport barriers. J. Atmos. Sci., 65, 855–874, https://doi.org/10.1175/2007JAS2227.1.
Eady, E. T., 1949: Long waves and cyclone waves. Tellus, 1, 33–52, https://doi.org/10.3402/tellusa.v1i3.8507.
Eden, C., 2007: Eddy length scales in the North Atlantic Ocean. J. Geophys. Res., 112, C06004, https://doi.org/10.1029/2006JC003901.
Eden, C., 2011: A closure for meso-scale eddy fluxes based on linear instability theory. Ocean Modell., 39, 362–369, https://doi.org/10.1016/j.ocemod.2011.05.009.
Eden, C., and R. J. Greatbatch, 2008: Towards a mesoscale eddy closure. Ocean Modell., 20, 223–239, https://doi.org/10.1016/j.ocemod.2007.09.002.
Ferrari, R., and M. Nikurashin, 2010: Suppression of eddy diffusivity across jets in the Southern Ocean. J. Phys. Oceanogr., 40, 1501–1519, https://doi.org/10.1175/2010JPO4278.1.
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, https://doi.org/10.1175/JPO2785.1.
Fox-Kemper, B., R. Ferrari, and R. Hallberg, 2008: Parameterization of mixed layer eddies. Part I: Theory and diagnosis. J. Phys. Oceanogr., 38, 1145–1165, https://doi.org/10.1175/2007JPO3792.1.
Fox-Kemper, B., and Coauthors, 2019: Challenges and prospects in ocean circulation models. Front. Mar. Sci., 6, 65, https://doi.org/10.3389/fmars.2019.00065.
Gent, P. R., and J. C. McWilliams, 1990: Isopycnal mixing in ocean circulation models. J. Phys. Oceanogr., 20, 150–155, https://doi.org/10.1175/1520-0485(1990)020<0150:IMIOCM>2.0.CO;2.
Gent, P. R., J. Willebrand, T. J. McDougall, and J. C. McWilliams, 1995: Parameterizing eddy-induced tracer transports in ocean circulation models. J. Phys. Oceanogr., 25, 463–474, https://doi.org/10.1175/1520-0485(1995)025<0463:PEITTI>2.0.CO;2.
Gnanadesikan, A., D. Bianchi, and M.-A. Pradal, 2013: Critical role for mesoscale eddy diffusion in supplying oxygen to hypoxic ocean waters. Geophys. Res. Lett., 40, 5194–5198, https://doi.org/10.1002/grl.50998.
Gnanadesikan, A., M.-A. Pradal, and R. Abernathey, 2015: Isopycnal mixing by mesoscale eddies significantly impacts oceanic anthropogenic carbon uptake. Geophys. Res. Lett., 42, 4249–4255, https://doi.org/10.1002/2015GL064100.
Gnanadesikan, A., A. Russell, M.-A. Pradal, and R. Abernathey, 2017: Impact of lateral mixing in the ocean on El Niño in a suite of fully coupled climate models. J. Adv. Model. Earth Syst., 9, 2493–2513, https://doi.org/10.1002/2017MS000917.
Green, J. S. A., 1960: A problem in baroclinic stability. Quart. J. Roy. Meteor. Soc., 86, 237–251, https://doi.org/10.1002/qj.49708636813.
Griesel, A., C. Eden, N. Koopmann, and E. Yulaeva, 2015: Comparing isopycnal eddy diffusivities in the Southern Ocean with predictions from linear theory. Ocean Modell., 94, 33–45, https://doi.org/10.1016/j.ocemod.2015.08.001.
Griffies, S. M., 1998: The Gent–McWilliams skew flux. J. Phys. Oceanogr., 28, 831–841, https://doi.org/10.1175/1520-0485(1998)028%3C0831:TGMSF%3E2.0.CO;2.
Groeskamp, S., S. M. Griffies, D. Iudicone, R. Marsh, A. J. Nurser, and J. D. Zika, 2019: The water mass transformation framework for ocean physics and biogeochemistry. Annu. Rev. Mar. Sci., 11, 271–305, https://doi.org/10.1146/annurev-marine-010318-095421.
Groeskamp, S., J. H. LaCasce, T. J. McDougall, and M. Rogé, 2020: Full-depth global estimates of ocean mesoscale eddy mixing from observations and theory. Geophys. Res. Lett., 47, e2020GL089425, https://doi.org/10.1029/2020GL089425.
Grooms, I., A. J. Majda, and K. S. Smith, 2015: Stochastic superparameterization in a quasigeostrophic model of the Antarctic Circumpolar Current. Ocean Modell., 85, 1–15, https://doi.org/10.1016/j.ocemod.2014.10.001.
Gruber, N., Z. Lachkar, H. Frenzel, P. Marchesiello, M. Münnich, J. C. McWilliams, T. Nagai, and G.-K. Plattner, 2011: Eddy-induced reduction of biological production in eastern boundary upwelling systems. Nat. Geosci., 4, 787–792, https://doi.org/10.1038/ngeo1273.
Hallberg, R., 2013: Using a resolution function to regulate parameterizations of oceanic mesoscale eddy effects. Ocean Modell., 72, 92–103, https://doi.org/10.1016/j.ocemod.2013.08.007.
Hogg, N. G., 2000: Low-frequency variability on the western flanks of the Grand Banks. J. Mar. Res., 58, 523–545, https://doi.org/10.1357/002224000321511007.
Holloway, G., 1986: Estimation of oceanic eddy transports from satellite altimetry. Nature, 323, 243–244, https://doi.org/10.1038/323243a0.
Holloway, G., and S. S. Kristmannsson, 1984: Stirring and transport of tracer fields by geostrophic turbulence. J. Fluid Mech., 141, 27–50, https://doi.org/10.1017/S0022112084000720.
Holmes, R. M., S. Groeskamp, K. D. Stewart, and T. J. McDougall, 2022: Sensitivity of a coarse-resolution global ocean model to a spatially variable neutral diffusivity. J. Adv. Model. Earth Syst., 14, e2021MS002914, https://doi.org/10.1029/2021MS002914.
Isachsen, P. E., 2011: Baroclinic instability and eddy tracer transport across sloping bottom topography: How well does a modified Eady model do in primitive equation simulations? Ocean Modell., 39, 183–199, https://doi.org/10.1016/j.ocemod.2010.09.007.
Isachsen, P. E., and O. A. Nøst, 2012: The air-sea transformation and residual overturning circulation within the Nordic Seas. J. Mar. Res., 70, 31–68, https://doi.org/10.1357/002224012800502372.
Jansen, M. F., A. J. Adcroft, R. Hallberg, and I. M. Held, 2015: Parameterization of eddy fluxes based on a mesoscale energy budget. Ocean Modell., 92, 28–41, https://doi.org/10.1016/j.ocemod.2015.05.007.
Jansen, M. F., A. J. Adcroft, S. Khani, and H. Kong, 2019: Toward an energetically consistent, resolution aware parameterization of ocean mesoscale eddies. J. Adv. Model. Earth Syst., 11, 2844–2860, https://doi.org/10.1029/2019MS001750.
Jones, C. S., and R. P. Abernathey, 2019: Isopycnal mixing controls deep ocean ventilation. Geophys. Res. Lett., 46, 13 144–13 151, https://doi.org/10.1029/2019GL085208.
Juricke, S., S. Danilov, N. Koldunov, M. Oliver, D. V. Sein, D. Sidorenko, and Q. Wang, 2020a: A kinematic kinetic energy backscatter parametrization: From implementation to global ocean simulations. J. Adv. Model. Earth Syst., 12, e2020MS002175, https://doi.org/10.1029/2020MS002175.
Juricke, S., S. Danilov, N. Koldunov, M. Oliver, and D. Sidorenko, 2020b: Ocean kinetic energy backscatter parametrization on unstructured grids: Impact on global eddy-permitting simulations. J. Adv. Model. Earth Syst., 12, e2019MS001855, https://doi.org/10.1029/2019MS001855.
Keffer, T., and G. Holloway, 1988: Estimating Southern Ocean eddy flux of heat and salt from satellite altimetry. Nature, 332, 624–626, https://doi.org/10.1038/332624a0.
Khani, S., M. F. Jansen, and A. Adcroft, 2019: Diagnosing subgrid mesoscale eddy fluxes with and without topography. J. Adv. Model. Earth Syst., 11, 3995–4015, https://doi.org/10.1029/2019MS001721.
Kjellsson, J., and L. Zanna, 2017: The impact of horizontal resolution on energy transfers in global ocean models. Fluids, 2, 45, https://doi.org/10.3390/fluids2030045.
Klocker, A., and R. Abernathey, 2014: Global patterns of mesoscale eddy properties and diffusivities. J. Phys. Oceanogr., 44, 1030–1046, https://doi.org/10.1175/JPO-D-13-0159.1.
Klocker, A., R. Ferrari, and J. H. LaCasce, 2012: Estimating suppression of eddy mixing by mean flows. J. Phys. Oceanogr., 42, 1566–1576, https://doi.org/10.1175/JPO-D-11-0205.1.
Kong, H., and M. F. Jansen, 2017: The eddy diffusivity in barotropic β-plane turbulence. Fluids, 2, 54, https://doi.org/10.3390/fluids2040054.
LaCasce, J. H., 2000: Floats and f/H. J. Mar. Res., 58, 61–95, https://doi.org/10.1357/002224000321511205.
LaCasce, J. H., 2008: Statistics from Lagrangian observations. Prog. Oceanogr., 77, 1–29, https://doi.org/10.1016/j.pocean.2008.02.002.
LaCasce, J. H., 2017: The prevalence of oceanic surface modes. Geophys. Res. Lett., 44, 11 097–11 105, https://doi.org/10.1002/2017GL075430.
LaCasce, J. H., and K. G. Speer, 1999: Lagrangian statistics in unforced barotropic flows. J. Mar. Res., 57, 245–274, https://doi.org/10.1357/002224099321618218.
LaCasce, J. H., and S. Groeskamp, 2020: Baroclinic modes over rough bathymetry and the surface deformation radius. J. Phys. Oceanogr., 50, 2835–2847, https://doi.org/10.1175/JPO-D-20-0055.1.
LaCasce, J. H., R. Ferrari, J. Marshall, R. Tulloch, D. Balwada, and K. Speer, 2014: Float-derived isopycnal diffusivities in the DIMES experiment. J. Phys. Oceanogr., 44, 764–780, https://doi.org/10.1175/JPO-D-13-0175.1.
Larichev, V. D., and I. M. Held, 1995: Eddy amplitudes and fluxes in a homogeneous model of fully developed baroclinic instability. J. Phys. Oceanogr., 25, 2285–2297, https://doi.org/10.1175/1520-0485(1995)025<2285:EAAFIA>2.0.CO;2.
Ledwell, J. R., A. J. Watson, and C. S. Law, 1993: Evidence for slow mixing across the pycnocline from an open-ocean tracer-release experiment. Nature, 364, 701–703, https://doi.org/10.1038/364701a0.
Ledwell, J. R., A. J. Watson, and C. S. Law, 1998: Mixing of a tracer in the pycnocline. J. Geophys. Res., 103, 21 499–21 529, https://doi.org/10.1029/98JC01738.
Lee, M.-M., A. J. G. Nurser, A. C. Coward, and B. A. de Cuevas, 2007: Eddy advective and diffusive transports of heat and salt in the Southern Ocean. J. Phys. Oceanogr., 37, 1376–1393, https://doi.org/10.1175/JPO3057.1.
Mak, J., D. P. Marshall, J. R. Maddison, and S. D. Bachman, 2017: Emergent eddy saturation from an energy constrained eddy parameterisation. Ocean Modell., 112, 125–138, https://doi.org/10.1016/j.ocemod.2017.02.007.
Mak, J., J. R. Maddison, D. P. Marshall, and D. R. Munday, 2018: Implementation of a geometrically informed and energetically constrained mesoscale eddy parameterization in an ocean circulation model. J. Phys. Oceanogr., 48, 2363–2382, https://doi.org/10.1175/JPO-D-18-0017.1.
Mak, J., A. Avdis, T. David, H. S. Lee, Y. Na, Y. Wang, and F. E. Yan, 2022a: On constraining the mesoscale eddy energy dissipation time-scale. J. Adv. Model. Earth Syst., 14, e2022MS003223, https://doi.org/10.1029/2022MS003223.
Mak, J., D. P. Marshall, G. Madec, and J. R. Maddison, 2022b: Acute sensitivity of global ocean circulation and heat content to eddy energy dissipation timescale. Geophys. Res. Lett., 49, e2021GL097259, https://doi.org/10.1029/2021GL097259.
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, https://doi.org/10.1175/JPO2949.1.
Martínez-Moreno, J., A. M. Hogg, and M. H. England, 2022: Climatology, seasonality, and trends of spatially coherent ocean eddies. J. Geophys. Res. Oceans, 127, e2021JC017453, https://doi.org/10.1029/2021JC017453.
McDougall, T. J., and P. C. McIntosh, 2001: The temporal-residual-mean velocity. Part II: Isopycnal interpretation and the tracer and momentum equations. J. Phys. Oceanogr., 31, 1222–1246, https://doi.org/10.1175/1520-0485(2001)031<1222:TTRMVP>2.0.CO;2.
Meredith, M. P., A. C. Naveira Garabato, A. M. Hogg, and R. Farneti, 2012: Sensitivity of the overturning circulation in the Southern Ocean to decadal changes in wind forcing. J. Climate, 25, 99–110, https://doi.org/10.1175/2011JCLI4204.1.
Nakamura, N., and D. Zhu, 2010a: Finite-amplitude wave activity and diffusive flux of potential vorticity in eddy–mean flow interaction. J. Atmos. Sci., 67, 2701–2716, https://doi.org/10.1175/2010JAS3432.1.
Nakamura, N., and D. Zhu, 2010b: Formation of jets through mixing and forcing of potential vorticity: Analysis and parameterization of beta-plane turbulence. J. Atmos. Sci., 67, 2717–2733, https://doi.org/10.1175/2009JAS3159.1.
Naveira Garabato, A. C., R. Ferrari, and K. L. Polzin, 2011: Eddy stirring in the Southern Ocean. J. Geophys. Res., 116, C09019, https://doi.org/10.1029/2010JC006818.
Nummelin, A., 2023a: Mesoscale eddies over sloping topography—Data. NIRD Research Data Archive, accessed 1 January 2024, https://doi.org/10.11582/2023.00129.
Nummelin, A., 2023b: Parameterizing mesoscale eddy buoyancy transport over sloping topography—Model setup. Zenodo, accessed 1 January 2024, https://doi.org/10.5281/zenodo.8227382.
Nummelin, A., and P. E. Isachsen, 2024: Parameterizing mesoscale eddy buoyancy transport over sloping topography. J. Adv. Model. Earth Syst., 16, e2023MS003806, https://doi.org/10.1029/2023MS003806.
Pedlosky, J., 1987: Geophysical Fluid Dynamics. 2nd ed. Springer-Verlag, 710 pp.
Pennel, R., and I. Kamenkovich, 2014: On the factors controlling the eddy-induced transport in the Antarctic Circumpolar Current. J. Phys. Oceanogr., 44, 2127–2138, https://doi.org/10.1175/JPO-D-13-0256.1.
Pradal, M.-A., and A. Gnanadesikan, 2014: How does the Redi parameter for mesoscale mixing impact global climate in an Earth System Model? J. Adv. Model. Earth Syst., 6, 586–601, https://doi.org/10.1002/2013MS000273.
Prandtl, L., 1925: Bericht über untersuchungen zur ausgebildeten turbulenz. Z. Angew. Math. Mech., 5, 136–139, https://doi.org/10.1002/zamm.19250050212.
Pringle, J. M., 2001: Cross-shelf eddy heat transport in a wind-free coastal ocean undergoing winter time cooling. J. Geophys. Res., 106, 2589–2604, https://doi.org/10.1029/2000JC900148.
Redi, M. H., 1982: Oceanic isopycnal mixing by coordinate rotation. J. Phys. Oceanogr., 12, 1154–1158, https://doi.org/10.1175/1520-0485(1982)012<1154:OIMBCR>2.0.CO;2.
Rhines, P., 1970: Edge-, bottom-, and Rossby waves in a rotating stratified fluid. Geophys. Fluid Dyn., 1, 273–302, https://doi.org/10.1080/03091927009365776.
Roach, C. J., D. Balwada, and K. Speer, 2016: Horizontal mixing in the Southern Ocean from Argo float trajectories. J. Geophys. Res. Oceans, 121, 5570–5586, https://doi.org/10.1002/2015JC011440.
Roach, C. J., D. Balwada, and K. Speer, 2018: Global observations of horizontal mixing from Argo float and surface drifter trajectories. J. Geophys. Res. Oceans, 123, 4560–4575, https://doi.org/10.1029/2018JC013750.
Sallée, J. B., K. Speer, and S. R. Rintoul, 2011: Mean-flow and topographic control on surface eddy-mixing in the Southern Ocean. J. Mar. Res., 69, 753–777, https://doi.org/10.1357/002224011799849408.
Seland, Ø., and Coauthors, 2020: Overview of the Norwegian Earth System Model (NorESM2) and key climate response of CMIP6 DECK, historical, and scenario simulations. Geosci. Model Dev., 13, 6165–6200, https://doi.org/10.5194/gmd-13-6165-2020.
Srinivasan, K., and W. R. Young, 2014: Reynolds stress and eddy diffusivity of β-plane shear flow. J. Atmos. Sci., 71, 2169–2185, https://doi.org/10.1175/JAS-D-13-0246.1.
Stammer, D., 1998: On eddy characteristics, eddy transports, and mean flow properties. J. Phys. Oceanogr., 28, 727–739, https://doi.org/10.1175/1520-0485(1998)028<0727:OECETA>2.0.CO;2.
Stewart, A. L., and A. F. Thompson, 2016: Eddy generation and jet formation via dense water outflows across the Antarctic continental slope. J. Phys. Oceanogr., 46, 3729–3750, https://doi.org/10.1175/JPO-D-16-0145.1.
Stewart, A. L., A. Klocker, and D. Menemenlis, 2018: Circum-Antarctic shoreward heat transport derived from an eddy- and tide-resolving simulation. Geophys. Res. Lett., 45, 834–845, https://doi.org/10.1002/2017GL075677.
Stewart, K. D., P. Spence, S. Waterman, J. Le Sommer, J.-M. Molines, J. M. Lilly, and M. H. England, 2015: Anisotropy of eddy variability in the global ocean. Ocean Modell., 95, 53–65, https://doi.org/10.1016/j.ocemod.2015.09.005.
Straub, D. N., 1994: Dispersive effects of zonally varying topography on quasigeostrophic Rossby waves. Geophys. Astrophys. Fluid Dyn., 75, 107–130, https://doi.org/10.1080/03091929408203650.
Taylor, G. I., 1921: Diffusion by continuous movements. Proc. London Math. Soc., 20, 196–212.
Thompson, A. F., 2010: Jet formation and evolution in baroclinic turbulence with simple topography. J. Phys. Oceanogr., 40, 257–278, https://doi.org/10.1175/2009JPO4218.1.
Tulloch, R., and Coauthors, 2014: Direct estimate of lateral eddy diffusivity upstream of Drake Passage. J. Phys. Oceanogr., 44, 2593–2616, https://doi.org/10.1175/JPO-D-13-0120.1.
Vallis, G. K., and M. E. Maltrud, 1993: Generation of mean flows and jets on a beta plane and over topography. J. Phys. Oceanogr., 23, 1346–1362, https://doi.org/10.1175/1520-0485(1993)023<1346:GOMFAJ>2.0.CO;2.
Wang, Y., and A. L. Stewart, 2020: Scalings for eddy buoyancy transfer across continental slopes under retrograde winds. Ocean Modell., 147, 101579, https://doi.org/10.1016/j.ocemod.2020.101579.
Wei, H., and Y. Wang, 2021: Full-depth scalings for isopycnal eddy mixing across continental slopes under upwelling-favorable winds. J. Adv. Model. Earth Syst., 13, e2021MS002498, https://doi.org/10.1029/2021MS002498.
Wei, H., Y. Wang, A. L. Stewart, and J. Mak, 2022: Scalings for eddy buoyancy fluxes across prograde shelf/slope fronts. J. Adv. Model. Earth Syst., 14, e2022MS003229, https://doi.org/10.1029/2022MS003229.
Wei, H., Y. Wang, and J. Mak, 2024: Parameterizing eddy buoyancy fluxes across prograde shelf/slope fronts using a slope-aware GEOMETRIC closure. J. Phys. Oceanogr., 54, 359–377, https://doi.org/10.1175/JPO-D-23-0152.1.
Wolfram, P. J., and T. D. Ringler, 2017: Quantifying residual, eddy, and mean flow effects on mixing in an idealized Circumpolar Current. J. Phys. Oceanogr., 47, 1897–1920, https://doi.org/10.1175/JPO-D-16-0101.1.
Wortham, C., and C. Wunsch, 2014: A multidimensional spectral description of ocean variability. J. Phys. Oceanogr., 44, 944–966, https://doi.org/10.1175/JPO-D-13-0113.1.
Wunsch, C., 2010: Toward a midlatitude ocean frequency-wavenumber spectral density and trend determination. J. Phys. Oceanogr., 40, 2264–2281, https://doi.org/10.1175/2010JPO4376.1.
Xie, J., H. Liu, and P. Lin, 2023: A multifaceted isoneutral eddy transport diagnostic framework and its application in the Southern Ocean. J. Adv. Model. Earth Syst., 15, e2023MS003728, https://doi.org/10.1029/2023MS003728.
Yankovsky, E., L. Zanna, and K. S. Smith, 2022: Influences of mesoscale ocean eddies on flow vertical structure in a resolution-based model hierarchy. J. Adv. Model. Earth Syst., 14, e2022MS003203, https://doi.org/10.1029/2022MS003203.
Zanna, L., P. Porta Mana, J. Anstey, T. David, and T. Bolton, 2017: Scale-aware deterministic and stochastic parametrizations of eddy-mean flow interaction. Ocean Modell., 111, 66–80, https://doi.org/10.1016/j.ocemod.2017.01.004.
Zika, J. D., J.-B. Sallée, A. J. S. Meijers, A. C. Naveira-Garabato, A. J. Watson, M.-J. Messias, and B. A. King, 2020: Tracking the spread of a passive tracer through Southern Ocean water masses. Ocean Sci., 16, 323–336, https://doi.org/10.5194/os-16-323-2020.