1. Introduction
The stratification and small aspect ratio of ocean mesoscale flows often makes them well approximated by two-dimensional (2D) dynamics. Even with the vast simplifications afforded by approximating mesoscale motions in this way, the emergent flow structures often bear excellent qualitative and dynamical resemblance to real ocean turbulence. Two-dimensional flows feature analogous cascades of enstrophy and energy as do quasigeostrophic (QG) flows, and in the presence of a Coriolis force bear similar flow structures such as vortices, jets, and filaments (Herring 1980; McWilliams 1984; Hua and Haidvogel 1986; McWilliams 1989). The generation of filaments has been of particular interest in the context of both two- and three-dimensional flows (e.g., Ashurst et al. 1987; Melander et al. 1987; Dritschel et al. 1991; Ohkitani and Kishiba 1995; Galanti et al. 1997; Kevlahan and Farge 1997; Von Hardenberg et al. 2000), and is manifested in the stretching and folding of fluid elements such that vorticity is accumulated into thin sets with sharp gradients. The accumulation of vorticity into these small-scale structures is the embodiment of the downscale enstrophy cascade (Chen et al. 2003), the study of which continues to draw interest due to its relevance to atmospheric and oceanic flows.
The enstrophy cascade itself, being a spectral quantity, is inherently nonlocal and thus challenging to examine in the highly heterogeneous flows of the real ocean. An alternative approach for studying its dynamics instead focuses on the processes governing the production of vorticity gradients, or equivalently the gradients of any tracer which obeys the same evolution equation as vorticity (McWilliams 1984; Weiss 1991; Ohkitani and Kishiba 1995; Protas et al. 1999; Straub 2003).
Regions where the production of gradients is particularly vigorous indicate where the turbulent cascades are especially active, and tend to lie outside the coherent structures associated with large-scale vortices (Mariotti et al. 1994). Methods used to identify vortices thus also implicitly indicate where gradient production is occurring, and by extension where the enstrophy cascade is strongest. These considerations have motivated a sizable body of literature dedicated to partitioning flows into regions that are dominated by strain (gradient-producing) versus those dominated by vorticity. The so-called “Okubo–Weiss parameter” (hereafter W; Okubo 1970; Weiss 1991) is perhaps the most well-known and easily applied criterion to partition ocean flows in this way, and has found many applications due to its facility with both model (e.g., Poje et al. 2010; Williams et al. 2011; Petersen et al. 2013) and altimetric data (e.g., Isern-Fontanet et al. 2003, 2004; Chelton et al. 2007; Henson and Thomas 2008; Souza et al. 2011, among others). It, too, was conceptualized for 2D flows, and its use in physical oceanography is justified largely by the quasi-horizontal nature of mesoscale motions.
Among oceanographers W is perhaps the most familiar solution to a more general problem in fluid dynamics, which is to develop a mathematical criterion with which to identify vortices. Several Eulerian schemes for vortex identification have been proposed in previous literature, many of which, like W, are based on an eigendecomposition of the velocity gradient tensor (e.g., Hunt et al. 1988; Chong et al. 1990; Berdahl and Thompson 1993; Zhou et al. 1999). Essentially all of these schemes are mathematically related to each other (Chakraborty et al. 2005); in fact, the Q criterion of Hunt et al. (1988) is identified as the three-dimensional generalization of W. Even more mathematically rigorous identification schemes have been developed which are invariant under rotations and translations (“objective”; e.g., Haller and Yuan 2000; Haller 2005; Haller et al. 2016), and which are shown to be superior to eigenvalue-based methods for vortex identification in unsteady and chaotic flows (e.g., Serra and Haller 2016; Pedergnana et al. 2020). Such methods are superior in their ability to track the material coherence of eddies, and recent work has shown that they can differ greatly from Eulerian methods in their estimates of ocean eddy transport (e.g., Abernathey and Haller 2018; Tarshish et al. 2018; Liu et al. 2019).
Nonetheless, among all of the available vortex identification schemes, the use of W remains prevalent in physical oceanography due to its simplicity and practicality. It is most commonly applied with altimetric data (e.g., Isern-Fontanet et al. 2003, 2004; Chelton et al. 2007, among others) to study mesoscale eddy characteristics, such as eddy lifetime, propagation distance, and size (e.g., Chelton et al. 2007, 2011). The prominence of mesoscale eddies and the strong tendency of their cores to be vorticity-dominated tends to make W-based detection methods reasonably skillful and quite visually compelling. The use of W as a diagnostic tool is not without shortcomings, however, and simple examples can be constructed where the criterion fails (e.g., Pierrehumbert and Yang 1993; Balmforth et al. 2000). In observational oceanography its primary deficiency, as described by Chelton et al. (2011), is that the geostrophic velocities that are used in its calculation are inferred from the sea surface height field h. The resulting equation for W is a quadratic function of the second derivatives of h, making it highly sensitive to noise in the altimetric data. For the particular application of eddy detection, W-based methods can also have difficulty distinguishing eddies from other strongly vortical features such as jet meanders (e.g., d’Ovidio et al. 2009; Souza et al. 2011; Williams et al. 2011). For these reasons eddy detection methods based on W are frequently employed using additional constraints, such as imposing a user-defined threshold to filter out regions where |W| is small and the flow structures are ambiguous (e.g., Isern-Fontanet et al. 2003; Morrow et al. 2004; Chelton et al. 2007).
This manuscript is partially motivated by the recognition that there are other vorticity-dominated features in the ocean that would not be properly classified as mesoscale eddies, but are nonetheless of dynamical interest. An obvious limitation of W as an ocean diagnostic is that it is purely a function of horizontal velocities and derivatives, whereas the ocean’s vertical shear and circulations drive some of the processes most fundamental to the broader climate system (e.g., Mahadevan and Tandon 2006; Thomas et al. 2008; Klein and Lapeyre 2009; Lévy et al. 2012). At large scales the velocity field is in approximate geostrophic and thermal wind balance, such that the vertical shears can be related to the horizontal derivatives of density via the QG approximation. This suggests an opportunity to explore new vortex identification schemes that have the same practical advantages as W (viz., that they can be diagnosed from surface fields via remote sensing) while still taking the vertical dimension into account. The increasing resolution of gridded observational products pertaining to h (e.g., Taburet et al. 2019; Zlotnicki et al. 2019), sea surface temperature (e.g., Donlon et al. 2012), and sea surface salinity (e.g., Droghei et al. 2018; Reul et al. 2020) means that observing structures smaller than mesoscale eddies is within reach (Reul et al. 2014; Umbert et al. 2015; Isern-Fontanet et al. 2016; Melnichenko et al. 2017; Vinogradova et al. 2019), and will become even more so with future remote sensing missions such as the Surface Water and Ocean Topography (SWOT; Fu et al. 2009), the proposed Copernicus Imaging Microwave Radiometer (CIMR; Kilic et al. 2018), and Soil Moisture and Ocean Salinity–High Resolution (SMOS-HR; Rodríguez-Fernández et al. 2019).
Looking beyond the application of vortex identification, the dynamics also provide motivation to explore new methods for separating strain- and rotation-dominated parts of the flow. Partitioning the flow in this way can yield information about which parts of the ocean favor vigorous turbulent cascades or are conduits for ocean ventilation (e.g., Klocker 2018; Bachman and Klocker 2020; Balwada et al. 2021), as well as their unique roles in the ocean energy cycle (e.g., Ferrari and Wunsch 2009, 2010; Storch et al. 2012; Chen et al. 2014). Part of the approach here is to take a critical look at the use of W in this context and to evaluate how it might fail, and alternatively, to consider whether any of the other previously mentioned schemes are better suited for exploring these topics. Last, the inclusion of the vertical dimension allows us to consider what geometric properties of these flow structures can be inferred from just the surface fields.
The purpose of this manuscript is to develop alternative Eulerian diagnostics to W that retain its appealing mathematical simplicity and applicability with surface diagnostics, while including dynamics associated with the vertical dimension. As such, the approach here begins by considering a QG flow, as opposed to the 2D framework typically used for W. Like many of the extant schemes mentioned above, the mathematics will involve an eigenanalysis based on decomposing the velocity gradient tensor, so that the results can be understood both algebraically (via eigenvalues) and geometrically (via eigenvectors).
The remainder of this paper is laid out as follows. Section 2 will review the mathematical concepts behind the Okubo–Weiss approach and why it becomes a degenerate mathematical problem in QG. An alternative approach based on the
2. The Okubo–Weiss approach for 2D and QG flows
a. The velocity gradient, strain-rate, and vorticity tensors
As this study will examine both two- and three-dimensional (QG) flows, this introductory discussion will establish the mathematical underpinnings of the Okubo–Weiss parameter in the 2D case before examining why this approach becomes degenerate in QG. Unlike the original derivations by Okubo (1970) and Weiss (1991), here the dynamics will be considered in the rotating frame to make the role of the Coriolis force explicit. For simplicity the motion will be considered on the f plane (which in the context of W is appropriate since it is entirely based on local velocity gradients) with constant Coriolis parameter f0, though it is possible to generalize these arguments to cases where f varies meridionally as well (Hua et al. 1998).
b. The 2D tracer gradient problem
The derivation of W in the 2D problem begins by either considering the evolution of particle trajectories (e.g., Okubo 1970) or tracer gradients (e.g., Weiss 1991), which are essentially dual approaches that yield the same overall interpretation (Hua and Klein 1998). Here the analysis will focus on the tracer gradient problem, and in particular we will consider a conserved tracer q. The choice to name this tracer q is done purposefully to evoke that these derivations also apply when considering the absolute vorticity (2D) or potential vorticity (QG) on the f plane, both of which are conserved in their respective flow regimes and are often denoted by q in the oceanographic literature.
A key point here is that both unique elements of the strain-rate tensor (σn and σs) appear in the growth rate for the gradient norm. The Okubo–Weiss parameter for 2D flows is thus consistent in the sense that it is completely described by the elements of
c. The degeneracy of the QG tracer gradient problem
That the eigenvalues of
3. The λ2 criterion for QG flows
Before discussing the merits of the
It is important to recognize that the family of schemes based on measuring the eigenvalues of
The scheme should somehow manifest the standard (2D) definition of W if uz = 0.
The parameters obtained through the scheme must be real, i.e., the eigenvalues of a symmetric tensor.
The small aspect ratio of QG flows means that the relevant strain and vorticity dynamics are quasi-horizontal. The tensor that yields these new parameters should have eigenvectors that reflect this quasi-horizontal geometry.
These criteria essentially insist that this scheme is treated as an extension of a 2D problem, in that purely 2D dynamics are recovered in the limit of no vertical shear. This section will introduce the
The
Before proceeding further it is worth examining the key features of (25) and whether they satisfy the desired bullet points at the beginning of this section. First, it is clear that this tensor is symmetric and its eigenvalues are thus guaranteed to be real. In the limit where uz = 0 the tensor as a whole reduces to the 2D version in (23), with zeros in the third row and column. It is thus perhaps unsurprising that, in similar fashion to (23), its eigenvalues reflect the influence of W. The final bullet point requires a derivation of the eigenvectors, and will make use of the asymmetry of the eigenvalues noted above.
a. The geometry of
For an arbitrary scalar field the Hessian matrix is a matrix of second derivatives that describes the curvature, or topology, of the field. Because the Hessian is symmetric its eigenvectors are orthogonal, and so its axes form a coordinate system with which to naturally describe the geometry of the field. In using the approach of the

Schematic showing the tilt of the principal axes of
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1

Schematic showing the tilt of the principal axes of
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1
Schematic showing the tilt of the principal axes of
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1
Summary of the eigendecomposition and geometry for each flow type.


b. A geometrically motivated extension of the Okubo–Weiss parameter
For real oceanic flows the
For the sake of simply identifying vortices the
Since no obvious choice for such a parameter emerges from the tensors or their geometry, one must construct the parameter using a combination of mathematical and physical insight. One possibility that was considered for W* was to use the determinant of
In all, this exploration of the
4. Simulations
Both ocean observations and numerical simulations are appropriate for testing the new parameters developed in this paper. Given that it is standard practice to use the Okubo–Weiss parameter, a fundamentally 2D quantity, for diagnoses even in 3D flows, we are not necessarily limited to using strictly 2D or QG flows for these tests. Here we choose to test the new parameters using a high-resolution primitive equation numerical simulation, whose output depicts a highly turbulent eddy field and contains dense information about both the velocity and buoyancy fields. The simulation output thus allows the new parameters to be diagnosed for many different types of flow structures, just as they would be applied for real ocean flows.
The simulation examined here is the same as was used in Bachman and Klocker (2020), whose key features are summarized below. The MITgcm (Marshall et al. 1997) was used to perform a 5-yr experiment of the Kerguelen Plateau region at 1/120° resolution (nominally 650-m grid spacing in each direction), with a domain extending from 60° to 85°E in the zonal direction and from 50° to 38°S in the meridional direction. The vertical grid consisted of 150 layers of varying thickness, ranging from 10 m at the surface to 50 m at depth. Open boundary conditions were used to force the model at the lateral boundaries, where the velocity, temperature, and salinity fields were forced using daily output from a larger simulation of the Indian Ocean sector of the Southern Ocean (Klocker 2018). A 1/2°-wide sponge layer was used to relax the model to the boundary conditions, with a one-day relaxation time scale at the inner edge of the sponge and a 4-h time scale at the outer edge of the sponge. The wind and buoyancy forcing was derived from annually and zonally averaged output from the Southern Ocean State Estimate (Mazloff et al. 2010) and applied as meridionally varying but constant-in-time forcing over the model domain.
Instantaneous snapshots of the velocities and buoyancy field were written out every six hours during the final year of simulated time. Given the very tight horizontal grid spacing, each snapshot resolved a vigorous and fast-evolving eddy field at nearly submesoscale-resolving resolution. While larger flow structures such as mesoscale eddies would be expected to exhibit QG dynamics, the submesoscale structures in the model feature Rossby numbers that are too large and balanced Richardson numbers that are too small for QG scaling to be appropriate (Thomas et al. 2008). The small-scale dynamics in the simulation are thus admittedly outside both the 2D and QG regimes that underpin the theory here. However, W and the

Visualizations of (a) surface buoyancy, (b) relative vorticity, (c) W, (d) W*, and the (e) first and (f) third eigenvalues of
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1

Visualizations of (a) surface buoyancy, (b) relative vorticity, (c) W, (d) W*, and the (e) first and (f) third eigenvalues of
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1
Visualizations of (a) surface buoyancy, (b) relative vorticity, (c) W, (d) W*, and the (e) first and (f) third eigenvalues of
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1
The middle and bottom rows of Fig. 2 show visualizations of the Okubo–Weiss parameter (labeled with a reminder that it is equal to 4 times the middle eigenvalue of
Here the reader is reminded that the sign of W* and W are the same at every point, so the main visual difference between Figs. 2c and 2d is due to baroclinicity that is accounted for by the |Q| terms in W*. Likewise, though all vortices in Fig. 2c still have negative values in Fig. 2d, the more baroclinic vortices stand out in Fig. 2d. Many of the vortices in Fig. 2d also appear as “doughnut” structures rather than disks because the baroclinicity is weakest at the vortex core. Last, note that the baroclinicity contributes significantly to the overall magnitude of the W* diagnostic, and as such the color scale in Fig. 2d is an order of magnitude larger than that of Fig. 2c. Figures 2e and 2f also are shown at this enlarged color scale, and confirm that eigenvalues

Histograms of the base-10 logarithm of W and W*, taken over all surface points in the model subdomain at year 5, day 100.
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1

Histograms of the base-10 logarithm of W and W*, taken over all surface points in the model subdomain at year 5, day 100.
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1
Histograms of the base-10 logarithm of W and W*, taken over all surface points in the model subdomain at year 5, day 100.
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1

Intensity plots showing the relationship between the magnitudes of F, ∇hb, W, and W* in logarithmic space. Values for these plots are gathered from surface points on year 5, day 100.
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1

Intensity plots showing the relationship between the magnitudes of F, ∇hb, W, and W* in logarithmic space. Values for these plots are gathered from surface points on year 5, day 100.
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1
Intensity plots showing the relationship between the magnitudes of F, ∇hb, W, and W* in logarithmic space. Values for these plots are gathered from surface points on year 5, day 100.
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1
The close relationship between F, |∇hb|, and W* is also clearly visible in spatial maps of these variables (Fig. 5), particularly with regard to strong frontal structures. Each panel of this figure shows values measured on year 5, day 100, where the range of the color axes are chosen to highlight how the tightest fronts and filaments are easily seen in all three metrics and match quite well between them. A particularly interesting result emerges by considering weaker values of these variables as well, where forming histograms of their magnitudes in logarithmic space reveals that all three have approximately lognormal distributions (Fig. 6). For the sake of identifying fronts by employing a threshold value on W*, this implies that a threshold based on standard deviations of log10|W*| would be likely to isolate the same strong features as would emerge by applying a threshold on the standard deviations of log10|F| or log10|∇hb|.

Spatial maps of (a) W*, (b) the frontogenesis function, and (c) the horizontal buoyancy gradient magnitude, highlighting their similarities for identifying strong fronts and filaments on year 5, day 100.
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1

Spatial maps of (a) W*, (b) the frontogenesis function, and (c) the horizontal buoyancy gradient magnitude, highlighting their similarities for identifying strong fronts and filaments on year 5, day 100.
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1
Spatial maps of (a) W*, (b) the frontogenesis function, and (c) the horizontal buoyancy gradient magnitude, highlighting their similarities for identifying strong fronts and filaments on year 5, day 100.
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1

Histograms of the magnitudes of F, ∇hb, and W* in logarithmic space, on year 5, day 100. Note that all three histograms show approximately lognormal distributions.
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1

Histograms of the magnitudes of F, ∇hb, and W* in logarithmic space, on year 5, day 100. Note that all three histograms show approximately lognormal distributions.
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1
Histograms of the magnitudes of F, ∇hb, and W* in logarithmic space, on year 5, day 100. Note that all three histograms show approximately lognormal distributions.
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1
Figure 7 shows the result of applying such a threshold to each of these variables. The vast majority of the features shown here are strain-dominated fronts, since they tend to be associated with larger values of |Q| than do vortices. In these plots an overbar represents the mean and σ represents the standard deviation of each variable’s magnitude in logarithmic space. Each panel shows the structures that are identified by taking only those points with magnitudes stronger than the mean plus half a standard deviation (first column), the mean plus one standard deviation (second column), and the mean plus 1.5 standard deviations (third column). Naturally, each column represents a progressively stronger filter moving from left to right across this figure (keeping approximately 31%, 16%, and 7% of points, respectively), which results in detecting fewer and thinner segments of these features. As seen previously in Figs. 6b–d, there is very close agreement between each variable in terms of which structures are detected, with differences generally pertaining to the precise size and shape of each feature. A noteworthy exception is highlighted in the green box in Fig. 7a, where a strongly vortical eddy (detected by all three metrics due to a strong temperature contrast between its fringe and its core) is detected quite clearly by the W* metric but slightly less so for the other two. For the stronger filters in the right two columns, more of this eddy is able to pass through the filter on W*, whereas it is effectively filtered out completely as measured by F and ∇hb. In practical terms these differences are unlikely to matter, since none of these variables are materially conserved and do not truly represent the stability of fluid trajectories in any case (e.g., Haller 2005); their similarity is presented here only to confirm the utility of W* for detecting strongly baroclinic structures.

Baroclinic structures detected by applying a threshold filter to the magnitudes of W*, F, ∇hb, and W in logarithmic space. The lognormal probability distributions for each variable (Fig. 6) suggest that these thresholds can be defined using standard deviations. Regions in white indicate where the magnitude is greater than the mean plus (left) 0.5, (center) 1.0, and (right) 1.5 standard deviations. The green box identifies an exceptionally baroclinic eddy. Values for these plots are gathered from surface points on year 5, day 300.
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1

Baroclinic structures detected by applying a threshold filter to the magnitudes of W*, F, ∇hb, and W in logarithmic space. The lognormal probability distributions for each variable (Fig. 6) suggest that these thresholds can be defined using standard deviations. Regions in white indicate where the magnitude is greater than the mean plus (left) 0.5, (center) 1.0, and (right) 1.5 standard deviations. The green box identifies an exceptionally baroclinic eddy. Values for these plots are gathered from surface points on year 5, day 300.
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1
Baroclinic structures detected by applying a threshold filter to the magnitudes of W*, F, ∇hb, and W in logarithmic space. The lognormal probability distributions for each variable (Fig. 6) suggest that these thresholds can be defined using standard deviations. Regions in white indicate where the magnitude is greater than the mean plus (left) 0.5, (center) 1.0, and (right) 1.5 standard deviations. The green box identifies an exceptionally baroclinic eddy. Values for these plots are gathered from surface points on year 5, day 300.
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1
To emphasize the superior skill of W* at identifying frontal structures, the bottom row of Fig. 7 shows the same standard deviation-based thresholds applied to log10|W|. In all three columns it is clear that W identifies an entirely different set of structures that are more circular in shape, which is indicative of its facility at identifying eddies in comparison to its identification of fronts (see also the discussion of Fig. 2).
Finally, Fig. 8 shows a map of the strain and vortex tilt as predicted by the new parameters, Π1 and Π3, respectively. The maps in Figs. 8b and 8d are filtered so that only regions where W > 0 and W < 0 are shown in color, respectively. Figures 8a and 8c show the frequency at which each value of Π1 and Π3 occurs in the map, where the vertical axis is in logarithmic scale. The convexity of the PDFs in Figs. 8a and 8c indicate that the distributions are not quite exponential, but are nonetheless dominated by values near −1 and 1, respectively. Likewise, Figs. 8b and 8d are dominated by darker shades, indicating values near |1|, except for some conspicuous regions of orange shading within certain vortices in Fig. 8d. As noted in section 3a, smaller values indicate where W is dominant over the contribution of

(a) Number of occurrences for each value of Π1 from (b) its spatial distribution throughout the model subdomain on year 5, day 100. (c),(d) Analogous results pertaining to Π3.
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1

(a) Number of occurrences for each value of Π1 from (b) its spatial distribution throughout the model subdomain on year 5, day 100. (c),(d) Analogous results pertaining to Π3.
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1
(a) Number of occurrences for each value of Π1 from (b) its spatial distribution throughout the model subdomain on year 5, day 100. (c),(d) Analogous results pertaining to Π3.
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-21-0037.1
Several other diagnostics were compared to the spatial maps of Π1 and Π3 to try to discern whether there exist correlations with other features of the flow. Among these were mixed layer depth, ratio of shear production to buoyancy flux (essentially to determine whether energy conversion in these eddies was more associated with lateral or vertical shears), and vortex size, but no firm correlations were found with any of these features. It is possible that Π1 and Π3 may be relatable to other, more complicated questions about ocean eddies—how do these parameters evolve over the lifetime of an eddy, for example, or is there any connection with the amount or distance of water mass transport—that are beyond the scope and capability of this dataset. It is also possible that Π1 and Π3 are simply mathematical novelties, given that they essentially represent the degree of structural tilt despite that the 3D QG equations have no vortex tilting term. Their precise relationship with the curvature of the pressure field could also be investigated, though it would require a specially designed QG model to do so since the steepest slopes of |1| are not resolvable with the aspect ratio,
5. Discussion and conclusions
A new Eulerian vortex detection scheme has been developed that aims to differentiate vorticity-dominated and strain-dominated regions in realistic ocean flows. Like other, earlier methods that are popularly used for eddy detection in observational and computational oceanography, the scheme differentiates these regions based on the velocity gradients of the flow. The motivation for the new scheme is the ocean’s quasi-two-dimensional behavior at large scales, which nonetheless features vertical shear and horizontal density gradients that are not accounted for by other methods that are based on the eigenvalues of the velocity gradient tensor. This issue stems from the fact that the nonzero eigenvalues of the 2D and QG velocity gradient tensors are identical, resulting in a mathematical degeneracy wherein the vertical shear is ignored. It is shown here, via pathological example, that the vertical shear can play a leading-order role in the evolution of tracer gradients in QG flow, thus necessitating a new scheme that accounts for the vertical dimension.
Under the QG approximation the vertical and horizontal dimensions are coupled through thermal wind balance. In principle this allows a scheme to be developed that is sensitive to dynamics in the vertical while still only measuring quantities in a horizontal plane, e.g., at the surface. In fact, since the QG tracer gradient problem is decoupled between the horizontal and vertical directions and the vertical shear only affects the vertical tracer gradient (e.g., Hua et al. 1998; Smith and Ferrari 2009), the information gained by measuring the vertical shear (the horizontal buoyancy gradient) is specific to the tracer evolution in the vertical. It is thus natural that a scheme that includes the horizontal buoyancy gradient would be sensitive those features where vertical dynamics are known to be significant, i.e., fronts and filaments.
The scheme developed here is based on the QG version of the
Vortex detection parameters are employed both for the analysis of remote sensing data and numerical model output, typically for estimating tracer transport by coherent eddies. A main advantage of 2D parameters like W is that they are naturally applied to data observed at the sea surface, which are readily available and densely sampled compared to observations at depth. The new parameter W* takes advantage of the predominantly geostrophic nature of large-scale flows that relates the vertical shear to horizontal density gradients via thermal wind balance, thus also being applicable to 2D (sea surface) flows despite including 3D dynamics. One of the primary advantages of W* is that it gives a window into vertical gradient dynamics despite only requiring 2D data. Many tracers considered to be important by oceanographers are either primarily forced at the surface (heat, via surface heat fluxes) or tend to collect near the surface (buoyant pollutants like oil, or chlorophyll), which would naturally establish vertical gradients. The evolution of the vertical gradients implied by large W* thus indicates important processes such as subduction or ventilation are occurring (i.e., the dynamics are stretching out or contracting the gradients, respectively), particularly at small-scale strain dominated regions as found by Balwada et al. (2021). Indeed, the new scheme has been shown to be superior at detecting precisely those features where subduction is known to strongly occur, such as fronts, filaments, and the fringes of coherent eddies (e.g., Stukel et al. 2017; Taylor et al. 2018; Freilich and Mahadevan 2020).
A subtle issue regarding how to apply W* concerns the kind of datasets for which it is most suitable. Parameters such as W are generally used only for vortex detection, not the detection of fronts, and are thus advantaged by the large [
The ability of W* to include baroclinicity in its measure of vortex strength is the primary way it is distinguished from W. Since by construction it has the same sign as W everywhere, it can essentially be considered a reweighted version of the Okubo–Weiss parameter that is particularly sensitive to nearly geostrophic, baroclinic features. This feature of W* is precisely what enables it to perform so well at identifying fronts and filaments, and may be useful for differentiating eddies based on their geometry (Fig. 8). However, for eddy detection it has the same limitations as W, namely, that it can be prone to misidentifying vorticity-dominated filaments as eddies (Fig. 2d). It also has the tendency to visually identify vortices as doughnuts rather than disks, since at the center of a vortex the |Q| term reaches its minimum, which may require automated vortex identification methods to be adjusted. Related to this point, baroclinicity causes W* to take on a far larger range of values than W, and one notable risk is that using a standard deviation-based threshold to distinguish eddies from the background flow may cause the scheme to miss the more weakly baroclinic vortices. For these reasons, users may find it more useful to simply use W for the vortex detection step, and then to employ W* as an auxiliary means of analysis or for the particular application of filament identification.
Finally, it is important to note the relative merits of Eulerian vortex identification methods such as W* versus those of Lagrangian methods (e.g., Beron-Vera et al. 2013; Froyland et al. 2015; Haller et al. 2016; Wang et al. 2016; Abernathey and Haller 2018). Several disadvantages of Eulerian methods are discussed by Haller (2015), most notably that they are not objective (i.e., the identified structures may differ depending on the rotation or translation of the reference frame) and are materially incoherent (subject to significant leakage through the identified boundaries of the structures). Eulerian methods thus tend to strongly overestimate the degree of material coherence of eddies, as well as the volume of fluid that remains trapped within the eddy core (Liu et al. 2019). For quantifying water mass transport by coherent eddies in large-scale or temporally filtered flows, Lagrangian methods are thus clearly superior. However, recent work by Sinha et al. (2019) demonstrated that Lagrangian methods are challenged at submesoscale- and internal gravity wave-resolving resolutions, where the appearance of intricate small-scale structures can obscure the large-scale transport barriers in the flow. They also found that high-frequency motions can lead to substantially higher leakage than would be detected if one instead used filtered velocities, meaning that Lagrangian methods might also overestimate material coherence when applied to observed (coarser-scale) velocity fields.
It is also important to note that the lifespan of an eddy does not necessarily coincide with the time that a bolus of water is trapped inside it. That is, eddies may persist even after leaking water, so Lagrangian methods may not be necessary for applications that do not require strict material coherence, such as tracking eddy lifetimes, propagation distance, or mechanisms of dissipation. Furthermore, Lagrangian methods are computationally very expensive, requiring high temporal resolution and millions of advected particles to realize their full potential. In comparison, Eulerian methods only require instantaneous snapshots of the flow field, and are cheap enough to be used over the entire globe for significant duration of time. They thus retain an important place in the study of ocean turbulence, and continue to be employed in the modern oceanographic literature (e.g., Faghmous et al. 2015; Cetina-Heredia et al. 2019). A novel parameter like W* that is capable of simultaneously detecting both eddies and fronts may open up new frontiers in how Eulerian methods are used, and future work will continue to explore this possibility.
Acknowledgments
This material is based upon work supported by the National Center for Atmospheric Research (NCAR), which is a major facility sponsored by the National Science Foundation under Cooperative Agreement 1852977.
Data availability statement
The numerical ocean model is available at http://mitgcm.org/. Additional data related to this paper may be requested from the author.
REFERENCES
Abernathey, R., and G. Haller, 2018: Transport by Lagrangian vortices in the eastern Pacific. J. Phys. Oceanogr., 48, 667–685, https://doi.org/10.1175/JPO-D-17-0102.1.
Ashurst, W. T., A. Kerstein, R. Kerr, and C. Gibson, 1987: Alignment of vorticity and scalar gradient with strain rate in simulated Navier–Stokes turbulence. Phys. Fluids, 30, 2343–2353, https://doi.org/10.1063/1.866513.
Bachman, S. D., and A. Klocker, 2020: Interaction of jets and submesoscale dynamics leads to rapid ocean ventilation. J. Phys. Oceanogr., 50, 2873–2883, https://doi.org/10.1175/JPO-D-20-0117.1.
Balmforth, N. J., W. R. Young, J. Fields, J.-L. Thiffeault, and C. Pasquero, 2000: Stirring and mixing: 1999 program of summer study in geophysical fluid dynamics. Tech. Rep. WHOI-2000-07, Woods Hole Oceanographic Institution, 302 pp., https://doi.org/10.1575/1912/94.
Balwada, D., Q. Xiao, S. Smith, R. Abernathey, and A. R. Gray, 2021: Vertical fluxes conditioned on vorticity and strain reveal submesoscale ventilation. J. Phys. Oceanogr., 51, 2883–2901, https://doi.org/10.1175/JPO-D-21-0016.1.
Basdevant, C., and T. Philipovitch, 1994: On the validity of the “Weiss criterion” in two-dimensional turbulence. Physica D, 73, 17–30, https://doi.org/10.1016/0167-2789(94)90222-4.
Berdahl, C., and D. Thompson, 1993: Eduction of swirling structure using the velocity gradient tensor. AIAA J., 31, 97–103, https://doi.org/10.2514/3.11324.
Beron-Vera, F. J., Y. Wang, M. J. Olascoaga, G. J. Goni, and G. Haller, 2013: Objective detection of oceanic eddies and the Agulhas leakage. J. Phys. Oceanogr., 43, 1426–1438, https://doi.org/10.1175/JPO-D-12-0171.1.
Bouali, M., O. Sato, and P. Polito, 2015: An algorithm to improve the detection of ocean fronts from whiskbroom scanner images. Remote Sens. Lett., 6, 942–951, https://doi.org/10.1080/2150704X.2015.1093187.
Cetina-Heredia, P., M. Roughan, E. van Sebille, S. Keating, and G. B. Brassington, 2019: Retention and leakage of water by mesoscale eddies in the East Australian Current system. J. Geophys. Res. Oceans, 124, 2485–2500, https://doi.org/10.1029/2018JC014482.
Chakraborty, P., S. Balachandar, and R. J. Adrian, 2005: On the relationships between local vortex identification schemes. J. Fluid Mech., 535, 189–214, https://doi.org/10.1017/S0022112005004726.
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, https://doi.org/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, https://doi.org/10.1016/j.pocean.2011.01.002.
Chen, R., G. R. Flierl, and C. Wunsch, 2014: A description of local and nonlocal eddy–mean flow interaction in a global eddy-permitting state estimate. J. Phys. Oceanogr., 44, 2336–2352, https://doi.org/10.1175/JPO-D-14-0009.1.
Chen, S., R. E. Ecke, G. L. Eyink, X. Wang, and Z. Xiao, 2003: Physical mechanism of the two-dimensional enstrophy cascade. Phys. Rev. Lett., 91, 214501, https://doi.org/10.1103/physrevlett.91.214501.
Chevillard, L., C. Meneveau, L. Biferale, and F. Toschi, 2008: Modeling the pressure Hessian and viscous Laplacian in turbulence: Comparisons with direct numerical simulation and implications on velocity gradient dynamics. Phys. Fluids, 20, 101504, https://doi.org/10.1063/1.3005832.
Chong, M. S., A. E. Perry, and B. J. Cantwell, 1990: A general classification of three-dimensional flow fields. Phys. Fluids A, 2, 765–777, https://doi.org/10.1063/1.857730.
Donlon, C. J., M. Martin, J. Stark, J. Roberts-Jones, E. Fiedler, and W. Wimmer, 2012: The operational sea surface temperature and sea ice analysis (OSTIA) system. Remote Sens. Environ., 116, 140–158, https://doi.org/10.1016/j.rse.2010.10.017.
d’Ovidio, F., J. Isern-Fontanet, C. López, E. Hernández-García, and E. García-Ladona, 2009: Comparison between Eulerian diagnostics and finite-size Lyapunov exponents computed from altimetry in the Algerian basin. Deep-Sea Res. I, 56, 15–31, https://doi.org/10.1016/j.dsr.2008.07.014.
Dritschel, D., P. Haynes, M. Juckes, and T. Shepherd, 1991: The stability of a two-dimensional vorticity filament under uniform strain. J. Fluid Mech., 230, 647–665, https://doi.org/10.1017/S0022112091000915.
Droghei, R., B. Buongiorno Nardelli, and R. Santoleri, 2018: A new global sea surface salinity and density dataset from multivariate observations (1993–2016). Front. Mar. Sci., 5, 84, https://doi.org/10.3389/fmars.2018.00084.
Elhmaïdi, D., A. Provenzale, and A. Babiano, 1993: Elementary topology of two-dimensional turbulence from a Lagrangian viewpoint and single-particle dispersion. J. Fluid Mech., 257, 533–558, https://doi.org/10.1017/S0022112093003192.
Faghmous, J. H., I. Frenger, Y. Yao, R. Warmka, A. Lindell, and V. Kumar, 2015: A daily global mesoscale ocean eddy dataset from satellite altimetry. Sci. Data, 2, 150028, https://doi.org/10.1038/sdata.2015.28.
Ferrari, R., and C. Wunsch, 2009: Ocean circulation kinetic energy: Reservoirs, sources, and sinks. Annu. Rev. Fluid Mech., 41, 253–282, https://doi.org/10.1146/annurev.fluid.40.111406.102139.
Ferrari, R., and C. Wunsch, 2010: The distribution of eddy kinetic and potential energies in the global ocean. Tellus, 62A, 92–108, https://doi.org/10.1111/j.1600-0870.2009.00432.x.
Freilich, M., and A. Mahadevan, 2020: Coherent pathways for subduction from the surface mixed layer at ocean fronts. J. Geophys. Res. Oceans, 126, e2020JC017042, https://doi.org/10.1029/2020JC017042.
Froyland, G., C. Horenkamp, V. Rossi, and E. Van Sebille, 2015: Studying an Agulhas ring’s long-term pathway and decay with finite-time coherent sets. Chaos, 25, 083119, https://doi.org/10.1063/1.4927830.
Fu, L.-L., D. Alsdorf, E. Rodriguez, R. Morrow, N. Mognard, J. Lambin, P. Vaze, and T. Lafon, 2009: The SWOT (Surface Water and Ocean Topography) mission: Spaceborne radar interferometry for oceanographic and hydrological applications. Proceedings of OceanObs’09: Sustained Ocean Observations and Information for Society, Vol. 2, J. Hall, D. E. Harrison, and D. Stammer, Eds., ESA Publication WPP-306, https://doi.org 10.5270/OceanObs09.cwp.33.
Galanti, B., J. Gibbon, and M. Heritage, 1997: Vorticity alignment results for the three-dimensional Euler and Navier-Stokes equations. Nonlinearity, 10, 1675–1694, https://doi.org/10.1088/0951-7715/10/6/013.
Haller, G., 2005: An objective definition of a vortex. J. Fluid Mech., 525, 1–26, https://doi.org/10.1017/S0022112004002526.
Haller, G., 2015: Lagrangian coherent structures. Annu. Rev. Fluid Mech., 47, 137–162, https://doi.org/10.1146/annurev-fluid-010313-141322.
Haller, G., and G. Yuan, 2000: Lagrangian coherent structures and mixing in two-dimensional turbulence. Physica D, 147, 352–370, https://doi.org/10.1016/S0167-2789(00)00142-1.
Haller, G., A. Hadjighasem, M. Farazmand, and F. Huhn, 2016: Defining coherent vortices objectively from the vorticity. J. Fluid Mech., 795, 136–173, https://doi.org/10.1017/jfm.2016.151.
Henson, S. A., and A. C. Thomas, 2008: A census of oceanic anticyclonic eddies in the Gulf of Alaska. Deep-Sea Res. I, 55, 163–176, https://doi.org/10.1016/j.dsr.2007.11.005.
Herring, J. R., 1980: Statistical theory of quasi-geostrophic turbulence. J. Atmos. Sci., 37, 969–977, https://doi.org/10.1175/1520-0469(1980)037<0969:RDOWTS>2.0.CO;2.
Hoskins, B. J., 1982: The mathematical theory of frontogenesis. Annu. Rev. Fluid Mech., 14, 131–151, https://doi.org/10.1146/annurev.fl.14.010182.001023.
Hoskins, B. J., I. Draghici, and H. Davies, 1978: A new look at the ω-equation. Quart. J. Roy. Meteor. Soc., 104, 31–38, https://doi.org/10.1002/qj.49710443903.
Hua, B. L., and D. B. Haidvogel, 1986: Numerical simulations of the vertical structure of quasi-geostrophic turbulence. J. Atmos. Sci., 43, 2923–2936, https://doi.org/10.1175/1520-0469(1986)043<2923:NSOTVS>2.0.CO;2.
Hua, B. L., and P. Klein, 1998: An exact criterion for the stirring properties of nearly two-dimensional turbulence. Physica D, 113, 98–110, https://doi.org/10.1016/S0167-2789(97)00143-7.
Hua, B. L., J. C. McWilliams, and P. Klein, 1998: Lagrangian accelerations in geostrophic turbulence. J. Fluid Mech., 366, 87–108, https://doi.org/10.1017/S0022112098001001.
Hunt, J. C., A. A. Wray, and P. Moin, 1988: Eddies, streams, and convergence zones in turbulent flows. Studying Turbulence Using Numerical Simulation Databases – II: Proceedings of the Summer Program 1988, Rep. CTR-S88, Center for Turbulence Researce, NASA, 193–208, http://web.stanford.edu/group/ctr/Summer/SP1988/19_HUNT.pdf.
Isern-Fontanet, J., E. García-Ladona, and J. Font, 2003: Identification of marine eddies from altimetric maps. J. Atmos. Oceanic Technol., 20, 772–778, https://doi.org/10.1175/1520-0426(2003)20<772:IOMEFA>2.0.CO;2.
Isern-Fontanet, J., J. Font, E. García-Ladona, M. Emelianov, C. Millot, and I. Taupier-Letage, 2004: Spatial structure of anticyclonic eddies in the Algerian basin (Mediterranean Sea) analyzed using the Okubo–Weiss parameter. Deep-Sea Res. II, 51, 3009–3028, https://doi.org/10.1016/j.dsr2.2004.09.013.
Isern-Fontanet, J., E. Olmedo, A. Turiel, J. Ballabrera-Poy, and E. García-Ladona, 2016: Retrieval of eddy dynamics from SMOS sea surface salinity measurements in the Algerian Basin (Mediterranean Sea). Geophys. Res. Lett., 43, 6427–6434, https://doi.org/10.1002/2016GL069595.
Jeong, J., and F. Hussain, 1995: On the identification of a vortex. J. Fluid Mech., 285, 69–94, https://doi.org/10.1017/S0022112095000462.
Kevlahan, N.-R., and M. Farge, 1997: Vorticity filaments in two-dimensional turbulence: Creation, stability and effect. J. Fluid Mech., 346, 49–76, https://doi.org/10.1017/S0022112097006113.
Kilic, L., and Coauthors, 2018: Expected performances of the Copernicus Imaging Microwave Radiometer (CIMR) for an all-weather and high spatial resolution estimation of ocean and sea ice parameters. J. Geophys. Res. Oceans, 123, 7564–7580, https://doi.org/10.1029/2018JC014408.
Klein, P., and G. Lapeyre, 2009: The oceanic vertical pump induced by mesoscale and submesoscale turbulence. Annu. Rev. Mar. Sci., 1, 351–375, https://doi.org/10.1146/annurev.marine.010908.163704.
Klein, P., B. L. Hua, and G. Lapeyre, 2000: Alignment of tracer gradient vectors in 2D turbulence. Physica D, 146, 246–260, https://doi.org/10.1016/S0167-2789(00)00119-6.
Klocker, A., 2018: Opening the window to the Southern Ocean: The role of jet dynamics. Sci. Adv., 4, eaao4719, https://doi.org/10.1126/sciadv.aao4719.
Lapeyre, G., P. Klein, and B. Hua, 1999: Does the tracer gradient vector align with the strain eigenvectors in 2D turbulence? Phys. Fluids, 11, 3729–3737, https://doi.org/10.1063/1.870234.
Lévy, M., R. Ferrari, P. J. Franks, A. P. Martin, and P. Rivière, 2012: Bringing physics to life at the submesoscale. Geophys. Res. Lett., 39, L14602, https://doi.org/10.1029/2012GL052756.
Liu, T., R. Abernathey, A. Sinha, and D. Chen, 2019: Quantifying Eulerian eddy leakiness in an idealized model. J. Geophys. Res. Oceans, 124, 8869–8886, https://doi.org/10.1029/2019JC015576.
Madrid, J. J., and A. M. Mancho, 2009: Distinguished trajectories in time dependent vector fields. Chaos, 19, 013111, https://doi.org/10.1063/1.3056050.
Mahadevan, A., and A. Tandon, 2006: An analysis of mechanisms for submesoscale vertical motion at ocean fronts. Ocean Modell., 14, 241–256, https://doi.org/10.1016/j.ocemod.2006.05.006.
Mancho, A. M., S. Wiggins, J. Curbelo, and C. Mendoza, 2013: Lagrangian descriptors: A method for revealing phase space structures of general time dependent dynamical systems. Commun. Nonlinear Sci. Numer. Simul., 18, 3530–3557, https://doi.org/10.1016/j.cnsns.2013.05.002.
Mariotti, A., B. Legras, and D. G. Dritschel, 1994: Vortex stripping and the erosion of coherent structures in two-dimensional flows. Phys. Fluids, 6, 3954–3962, https://doi.org/10.1063/1.868385.
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, 5753–5766, https://doi.org/10.1029/96JC02775.
Mazloff, M. R., P. Heimbach, and C. Wunsch, 2010: An eddy-permitting southern ocean state estimate. J. Phys. Oceanogr., 40, 880–899, https://doi.org/10.1175/2009JPO4236.1.
McWilliams, J. C., 1984: The emergence of isolated, coherent vortices in turbulent flow. AIP Conf. Proc., 106, 205–221, https://doi.org/10.1063/1.34273.
McWilliams, J. C., 1989: Statistical properties of decaying geostrophic turbulence. J. Fluid Mech., 198, 199–230, https://doi.org/10.1017/S0022112089000108.
McWilliams, J. C., 2019: A survey of submesoscale currents. Geosci. Lett., 6, 3, https://doi.org/10.1186/s40562-019-0133-3.
Melander, M., J. McWilliams, and N. Zabusky, 1987: Axisymmetrization and vorticity-gradient intensification of an isolated two-dimensional vortex through filamentation. J. Fluid Mech., 178, 137–159, https://doi.org/10.1017/S0022112087001150.
Melnichenko, O., A. Amores, N. Maximenko, P. Hacker, and J. Potemra, 2017: Signature of mesoscale eddies in satellite sea surface salinity data. J. Geophys. Res. Oceans, 122, 1416–1424, https://doi.org/10.1002/2016JC012420.
Morrow, R., F. Birol, D. Griffin, and J. Sudre, 2004: Divergent pathways of cyclonic and anti-cyclonic ocean eddies. Geophys. Res. Lett., 31, L24311, https://doi.org/10.1029/2004GL020974.
Nencioli, F., C. Dong, T. Dickey, L. Washburn, and J. C. McWilliams, 2010: A vector geometry–based eddy detection algorithm and its application to a high-resolution numerical model product and high-frequency radar surface velocities in the Southern California Bight. J. Atmos. Oceanic Technol., 27, 564–579, https://doi.org/10.1175/2009JTECHO725.1.
Ohkitani, K., and S. Kishiba, 1995: Nonlocal nature of vortex stretching in an inviscid fluid. Phys. Fluids, 7, 411–421, https://doi.org/10.1063/1.868638.
Okubo, A., 1970: Horizontal dispersion of floatable particles in the vicinity of velocity singularities such as convergences. Deep Sea Res. Oceanogr. Abstr., 17, 445–454, https://doi.org/10.1016/0011-7471(70)90059-8.
Pedergnana, T., D. Oettinger, G. P. Langlois, and G. Haller, 2020: Explicit unsteady Navier–Stokes solutions and their analysis via local vortex criteria. Phys. Fluids, 32, 046603, https://doi.org/10.1063/5.0003245.