1. Introduction
General circulation models (GCMs) are limited in their ability to resolve the gordian interactions between the atmosphere, ocean, land, and biology over all relevant time and spatial scales. Hence, they simulate these interactions by directly including as many processes as computational constraints permit, while processes too small, fast, or complex are approximated through parameterizations. Parameterizations are reduced mathematical expressions to capture the dominant impacts, while remaining computationally efficient and complementary with other components of the GCM.
The ocean surface layer is the most turbulent layer in the ocean, driven primarily by winds, waves and buoyancy forcing. The ocean surface layer contains the mixed layer, which can be described broadly as the layer in which temperature, salinity, and other tracers are vertically well mixed. It is the connecting layer with the atmospheric boundary layer, where air–sea fluxes take place, and links the deep, stratified ocean waters with the free atmosphere through vertical mixing and surface ventilation (Fox-Kemper et al. 2022). Furthermore, vertical mixing processes near the ocean surface are critical in transporting tracers and supplying essential nutrients to marine biology, as the euphotic zone where primary productivity occurs is coincident with the surface layer (Taylor and Ferrari 2011; Smith et al. 2016; Mahadevan 2016; Olita et al. 2017; Lévy et al. 2018).
Submesoscales span the range of 0.1–10 km in horizontal scale, 0.01–1 km in vertical scale, and from hours to days in time evolution, and boundary layer turbulence is on the order of meters in length and from minutes to hours in time (e.g., Grant and Belcher 2009; McWilliams 2016). Therefore, submesoscales and boundary layer turbulence tend to be on scales smaller than the grid used in GCMs, even at the highest possible resolution (Dong et al. 2020, 2021). Accurate representation of these processes is especially important for climate simulations as the mixing-versus-restratification balance near the surface determines the effective mixed layer depth, which is an effective marker of overall climate sensitivity in models (Li et al. 2019; Fox-Kemper et al. 2019; Hall and Fox-Kemper 2021, manuscript submitted to Geophys. Res. Lett.).
The restratification process (i.e., mixed layer shoaling) associated with submesoscales is represented by the mixed layer eddy (MLE) parameterization (Fox-Kemper et al. 2011), which captures the effects of baroclinic mixed layer eddies that form along submesoscale fronts in a weakly stratified background state, that is, the mixed layer (Nurser and Zhang 2000; Boccaletti et al. 2007; Fox-Kemper et al. 2008).
Note that the MLE parameterization only intends to provide a vertical flux
The substitution of
Dimensional parameters. Note that we do not need any of the thermal expansion parameters β or g, because they are represented by
However, over the past decade, several studies have shown these assumptions to be overly restrictive on the one hand, as argued by Calvert et al. (2020) that due to a natural cancellation of the NH/f formulation with the other buoyancy gradients in (6) and (3), the cutoff Lf,min is not needed under careful numerical implementation. On other hand, this scaling has been shown to be too simplistic, especially in the presence of surface forcing, such as winds and convection (e.g., Mahadevan et al. 2010; Callies and Ferrari 2018a) or in simulations with initial frontal widths selected far from this value or freely evolving without mesoscale strain (Callies and Ferrari 2018b). This uncertainty highlights the strong dependence of the parameterization (7) on frontal width Lf, for which no mechanistic scaling law currently exists to comprehensively evaluate it, or its relationship to the turbulent processes that help set it.
The process involving the formation and sharpening of fronts is known as frontogenesis (e.g., Hoskins and Bretherton 1972; McWilliams 2021). In the ocean, the mixed layer is weakly stratified in the vertical, thus horizontal density gradients can become dominant, and strengthen to form sharp horizontal fronts. Two primary mechanisms are found to onset frontogenesis: a density gradient in the presence of (i) an external strain field (e.g., mesoscale eddies), and (ii) vertical turbulent fluxes (e.g., boundary layer turbulence).
Classic strain-induced frontogenesis theory was originally developed by Hoskins and Bretherton (1972) in a semigeostrophic framework, which assumes geostrophic balance in the alongfront direction, thus reducing to a 2D inviscid, adiabatic flow (i.e., no turbulence) in the cross-frontal plane. Classic frontogenesis theory is able to describe frontal dynamics at leading order, by solving for the alongfront current (at the top and bottom of the mixed layer) and a cross-frontal ageostrophic overturning circulation. However, it results in an infinitely narrow front in a finite time, an unphysical limit that does not comply with observations (Hosegood et al. 2006; Ramachandran et al. 2018; Johnson et al. 2020) or numerical simulations (Suzuki et al. 2016; Sullivan and McWilliams 2018). Furthermore, semigeostrophy is found to be inaccurate for 3D flows such as elongated fronts where curvature matters or during late frontogenesis when instabilities develop along the front (Rotunno et al. 1994; Gent et al. 1994).
The process of turbulence-induced frontogenesis critically involves the turbulent thermal wind (TTW) balance developed by Gula et al. (2014) and McWilliams et al. (2015), also analogous to the generalized Ekman equation (Cronin and Kessler 2009). TTW theory is found to be more consistent with ocean submesoscales, allowing for larger Rossby numbers as fronts and filaments arrive at scales much sharper than described by semigeostrophic theory (McWilliams 2021). However, it is very difficult to observe the strain field and turbulent fluxes in the ocean, and observations of TTW are scarce. Regional ocean models have been useful tools for studying TTW behavior in a near-realistic environment (e.g., Gula et al. 2014; Dauhajre and McWilliams 2018), and only recently have high-resolution numerical simulations been able to simulate the multiscale TTW range. Sullivan and McWilliams (2018, 2019) are among the first to simulate dense filaments undergoing TTW frontogenesis in the presence of realistic surface boundary layer turbulence.
Both strain-induced and turbulence-induced frontogenesis proceed until disrupted by turbulent fluxes that arrest the ever strengthening front (Bodner et al. 2020). It is also possible for strain-induced fronts to be advected out of the frontogenetic confluent strain region and thereby stop frontogenesis before the occurrence of arrest (or singularity). These turbulent fluxes can result from frontal instabilities or ocean surface forcing such as winds, waves, and surface cooling (McWilliams 2021).
There are various instabilities that populate the upper ocean, due to the effects of winds, waves and stratification, that may affect the evolution of fronts: gravitational instability or pure convection (Haine and Marshall 1998); symmetric instability (SI), which mixes along isopycnals in the presence of down-front winds and negative potential vorticity (PV) (Hoskins 1974; Thomas et al. 2013; Bachman et al. 2017); mixed layer baroclinic instability, which acts to restratify the mixed layer by slumping buoyancy gradients from horizontal to vertical (Boccaletti et al. 2007; McWilliams et al. 2009; McWilliams and Molemaker 2011); Langmuir turbulence, which creates convergence zones at the ocean surface and contributes to upper ocean mixing (McWilliams et al. 1997; Hamlington et al. 2014; Suzuki et al. 2016); horizontal shear instability due to the sharpening front itself (Sullivan and McWilliams 2018, 2019); and other mixing, wave breaking, and topographic effects (Garrett and Loder 1981; Thompson 2000; Teixeira and Belcher 2002; Nagai et al. 2006; Sullivan et al. 2007; Gula et al. 2016). Some of these phenomena require horizontal gradients such as a front to exist (e.g., symmetric instability or baroclinic instability), while others are related more generally to surface forcing (e.g., boundary layer turbulence).
Each of these instabilities may be recognized by characteristic energy sources, scale, and dependence on favorable stratification or shear conditions (Haney et al. 2015). In the ocean, observations are rarely able to simultaneously and conclusively isolate this set of constraints, so novel theoretical and modeling approaches are useful to study these processes and how they interact in a more idealized setting. Submesoscale and boundary layer turbulence can be differentiated by energetic properties (e.g., Hosegood et al. 2006; Haney et al. 2015), lack of hydrostasy (Hamlington et al. 2014) and dynamical markers such as PV (Bodner and Fox-Kemper 2020). Large-eddy simulations (LES) are particularly useful for their study. As computational capabilities have increased in recent years, several studies have modeled the multiscale interactions between submesoscales and boundary layer turbulence.
Skyllingstad and Samelson (2012) studied the interaction between MLE and small-scale turbulence using a nonhydrostatic LES of a warm filament in the presence of Langmuir turbulence, focusing primarily on the transfer of energy between MLE and boundary layer turbulence. Hamlington et al. (2014) studied the weak interaction limit between boundary layer turbulence and submesoscale eddies, and the associated instabilities that arise from this interaction. They compare cases of shear turbulence driven by wind stress, with Langmuir turbulence driven by wind and wave effects. Haney et al. (2015) focus more specifically on how wave effects can alter the PV field and promote certain instabilities along a submesoscale front. Suzuki et al. (2016) identified a strain-induced front in a subdomain from the Hamlington et al. (2014) LES and investigates what energizes and torques the submesoscale front in the presence of waves. Crowe and Taylor (2020) study the evolution of an idealized TTW submesoscale front under varying wind stress and buoyancy flux conditions.
Few studies on interactions between submesoscales and boundary layer turbulence focus on the mechanism of frontogenetic arrest and how it selects stable frontal width. Bodner et al. (2020) propose a theoretical framework for the effects of turbulence on frontal formation in the quasigeostrophic limit. Vertical turbulent fluxes were found to enhance frontogenesis whereas horizontal fluxes are able to oppose it. The tendency and effects of turbulent processes on frontogenesis in Bodner et al. (2020) are consistent with LES studies of frontal evolution (e.g., McWilliams 2017; Sullivan and McWilliams 2018, 2019), where vertical turbulent fluxes assist frontogenesis, and horizontal processes can arrest or contribute to its decay. Note that both vertical and horizontal instabilities on small scales will lead to isotropic 3D turbulence and serve both roles eventually.
These studies have set the scene for a more physical estimate of frontal width, accounting for the interactions between fronts and boundary layer turbulence to determine a scale comparable to that observed. Here we propose and test a new scaling motivated by these principles. In section 2 the new frontal width scaling law is presented that relates Lf to surface forcing parameters. Based on TTW theory, this scaling highlights key balances required for frontogenesis and frontogenetic arrest. Section 3 tests the new scaling in a collection of large-eddy simulations and data from a realistically forced General Ocean Turbulence Model (GOTM) ensemble. Some impacts of the new parameterization as implemented in CESM2 are discussed in section 4. Summary and discussion are given in section 5.
2. A new scaling for frontal width
We next use scale analysis to explore this equation and the horizontal scale that sets this three-way balance.
Dimensionless parameters.
In cases forced with both convection and stress, boundary layer mixing schemes such as the K-profile parameterization (KPP; Large et al. 1994) and the energetics-based planetary boundary layer scheme (ePBL; Reichl and Hallberg 2018) tend to use combinations of the convective and friction velocities. A convenient formulation from ePBL’s energy budget considerations is
For a three-way balance to hold, the following conditions need to be met,
- 1) Ri2Ro2 = O(1),
- 2) Ek2 = O(1), and
- 3) Ri2Ro2/Ek2 = O(1).
These conditions select for a submesoscale front, where both boundary layer turbulence theory and geostrophic theory balance simultaneously, which together are the essence of TTW theory. In the submesoscale, the Rossby and Richardson numbers are O(1).
From the second condition, h emerges as proportional to the Ekman depth, that is,
This new formula on the right of (27) has a number of immediately apparent advantages over its predecessor on the left. The ad hoc renormalization mixing time scale τ is not needed because the f factors cancel from the denominator. This is an important improvement not only because it eliminates an artificial parameter, but also because the new parameterization naturally vanishes at the equator (Ψ → 0 as f → 0), where the submesoscales become resolved (e.g., Dong et al. 2020). Furthermore, Lf no longer includes a determination of the mixed layer stratification N, which is not robustly extracted from the effects of many boundary layer vertical mixing parameterizations, especially bulk boundary layer formulations (e.g., Kraus and Turner 1967; Price et al. 1986). Interestingly, the cancellation exploited for numerical robustness in the face of uncertain Lf by Calvert et al. (2020) does not occur in this formulation. In the limit where the eddies are given sufficient time to enlarge and their nonlinear scale width grows to exceed the frontal width, the scaling (4) appears to underestimate the restratification rate (Callies and Ferrari 2018b). Including the smaller of the two scales—eddy and front—in the denominator of (27) tends toward reducing this bias, although in this regime the eddy scale is not known (as it is involved in an inverse cascade from initiation at the linear instability scale and thus depends on the unknown time since the onset of instability). In the opposite case, where the front is wider than the instability scale, (27) produces a corrected average frontal strength as argued by Fox-Kemper et al. (2011). It is also important to emphasize that this new formula now involves both the boundary layer depth h and the mixed layer depth H, which are two distinct quantities in climate models. Finally, as was the objective, the surface forcing parameters
3. Proofs of concept
a. Testing in LES
A suite of high-resolution LES are presented and analyzed, where they capture the multiscale character of strain-induced frontogenesis, as it interacts with different instability and mixing mechanisms such as wind forcing and waves, mixed layer instabilities, convection and symmetric instability. In all runs, after instabilities have saturated, the cross-frontal scale halts at a constant width and does not become infinitesimally thin, as theory would predict. Sullivan and McWilliams (2018, 2019) conducted similar experiments for a range of LES, simulating a filament with varying surface forcing conditions. Specifically, they examined the process of frontal formation, arrest and decay of a turbulent-induced (TTW) cold filament in the presence of wind stress, convection and waves. Further details on how these simulations are set up and analyzed can be found in the supplemental material.
We utilize results from these two types of LES (e.g., turbulence-induced and strain-induced frontogenesis) to test the new Lf under various surface forcing conditions (
Following the SM setup, the coefficients in (28) are those for aligned winds and waves and the Langmuir number is La = 0.32. Note that the enhancement factor is only applied to
Horizontal shear instability is found to be associated with frontogenetic arrest, where Ri ≤ 0.25 in most strain-induced cases in the supplemental material and as discussed in Sullivan and McWilliams (2018, 2019) and Bodner et al. (2020). Note that it is assumed that the arresting turbulence is isotropic, thus horizontal eddy viscosity is equivalent to vertical eddy viscosity. Hence, we set the Richardson number in the frontal width scaling to match the assumed frontogenetic arrest physics of stratified shear turbulence: Ri = 0.25. Furthermore, as shown in Table 3 the constants c1, c2, and c3 are confirmed by direct diagnosis to be of order 1 in the strain-induced LES. Thus, we choose to set the constant CL = Ri = 0.25. With a more extensive set of arrested-front LES, this parameter estimate could be improved upon, but each of the LES must span from the shear turbulence scale through to the submesoscale, so they are costly and thus an ensemble of convenience was analyzed to suit the purpose approximately.
Estimate of parameterization constants given from strain-induced runs.
Figure 1 is a log–log plot of the measured
Log–log plot of the measured
Citation: Journal of Physical Oceanography 53, 1; 10.1175/JPO-D-21-0297.1
b. Estimates from the GOTM
The latest extension of the GOTM extended to incorporate the CVmix GCM mixing parameterizations (Levy et al. 2014; Li et al. 2021) may be used as a stand-alone model for studying dynamics of boundary layers in natural waters (e.g., Umlauf and Burchard 2005; Li et al. 2019). Here we run GOTM with observed forcing and initialization to provide h,
The histograms corresponding to values of Lf in the four regions are shown in Fig. 2. Although seasonality and monsoon variability will impact this distribution beyond the window of time from the observations, the results are in the range of the submesoscale frontal widths observed, generally O(1) km, with occasionally sharper or wider fronts (e.g., Pallàs-Sanz et al. 2010; Thompson et al. 2016; Ramachandran et al. 2018). Nonetheless, a more quantitative evaluation of the scaling compared with observations is still needed.
The Lf (m) estimated from
Citation: Journal of Physical Oceanography 53, 1; 10.1175/JPO-D-21-0297.1
c. Evaluations on a global scale
To obtain global statistics of the new Lf scaling we take advantage of the unique configuration from Dong et al. (2020), which uses the recent submesoscale-permitting MITgcm-LLC4320 in tandem with GOTM (see also Rocha et al. 2016; Su et al. 2018). Ocean state variables and surface fluxes (e.g.,
Global maps of the old Lf scaling during summer and winter are shown in the upper panels of Fig. 3 and are compared with global maps of the new Lf scaling in the middle panels and zonal median in the bottom panels. The new Lf in Fig. 3 exhibits values ranging from roughly 50 km at the equator to 10 m in the high latitudes, accounting for much sharper fronts than previously obtained by the deformation radius estimate, with a minimum value of only 1 km. This is not to assume that all fronts at high latitudes are 10 m wide, but rather that the contribution of unresolved fronts is given on average from sharper fronts via the k−2 spectral slope estimate (Fox-Kemper et al. 2011). In the following section we examine aspects of sensitivity in climate models due to this change in scale factor.
Global maps of Lf calculated using the (a),(b) old scaling based on the deformation radius NH/f, (c),(d) new scaling based on (24), and (e),(f) their zonal median, with shaded regions denoting the 10th and 90th percentiles. Global winter is given from February and August in the Northern and Southern Hemispheres, respectively, with the opposite during summer. Note that we exclude values of old Lf within 5° of the equator as they become exponentially large.
Citation: Journal of Physical Oceanography 53, 1; 10.1175/JPO-D-21-0297.1
4. Implementation
The new MLE parameterization formula (27) was implemented in the CESM2.1.3-POP model, where the newly required parameters
The new parameterization was tested in two types of simulations: a global fully coupled simulation and a CORE-v2 forced ocean and sea ice simulation, which covers forcing data from 1948 to 2009 (Large and Yeager 2009). The fully coupled model was run for 100 years for both the new and control versions of the MLE parameterizations. The forced model was run for five cycles amounting to a total of 310 simulated years. The mixed layer depth climatology was obtained by averaging over the last 20 years of the coupled simulation and last cycle in the forced simulations.
Figure 4 helps visualize the global values of new Lf across the different CESM simulations. Summer and winter results in both the coupled and forced simulations generally resemble the new Lf values shown in Fig. 3. In particular, the zonal median of new and old Lf shown in Figs. 4e and 4f truly highlight the artificial nature of the old Lf and the imposed Lf,min as compared with the sharper fronts obtained by the new Lf scaling.
As in Fig. 3, but for global maps of the new Lf scaling calculated from CESM (a),(b) coupled and (c),(d) forced simulations. (e),(f) Their zonal median is compared with the old Lf based on (7) with Lf,min = 5 km (thin lines) used by the control simulations. Here global winter is given from the average of January–March in the Northern Hemisphere and July–September in the Southern Hemisphere, with the opposite during summer.
Citation: Journal of Physical Oceanography 53, 1; 10.1175/JPO-D-21-0297.1
Figures 5 and 6 show the mixed layer depth during summer and winter in both hemispheres, respectively, for observations given by the de Boyer Montégut et al. (2004) dataset updated to include Argo data up to 2012, and compared with the control and new parameterization for both the coupled and forced simulations. Most importantly, the results appear to be qualitatively similar in all simulations, and generally resemble observations. This is demonstrated in the top-right panels of Figs. 5 and 6 where scatterplots of the MLD given by new Lf versus control in all simulations and over all grid points are on the same order.
Global maps of mean mixed layer depth during summer in the Northern Hemisphere (July–September) and Southern Hemisphere (January–March) from (a) observations given by the de Boyer Montégut et al. (2004) dataset updated to include Argo data up to 2012, (b),(c) MLD given by new Lf vs control in all simulations and over all grid points; (d),(e) corresponding global maps in the coupled simulations and (f),(g) forced simulations.
Citation: Journal of Physical Oceanography 53, 1; 10.1175/JPO-D-21-0297.1
As in Fig. 5, but during winter in the Northern Hemisphere (January–March) and Southern Hemisphere (July–September).
Citation: Journal of Physical Oceanography 53, 1; 10.1175/JPO-D-21-0297.1
Some sensitivity of the MLD to the new parameterization is apparent and the parts in the world that exhibit these changes are climatically important: for example, the tropical and equatorial summertime regions and near sea ice-covered regions (see observations and discussion in Timmermans and Winsor 2013), such as south of 60°S in austral wintertime where Antarctic Bottom Water forms and the subpolar Atlantic and Greenland and Icelandic Seas. The differences in MLD between the control and new parameterization are highlighted in Figs. 7 and 8 for the coupled and forced simulations, respectively. Multiple panels with custom color bar ranges help identify MLD adjustments on regional scales. These reveal that in the coupled simulations, the new parameterization leads to MLD changes in the North Atlantic, deepening in the Southern Ocean, and shoaling in the equator and Indian Ocean, where reducing the bias has been difficult to achieve by altering boundary layer mixing schemes alone (e.g., Li et al. 2019). Interestingly, although similar global patterns are also apparent in the forced simulations, no significant reduction in the equatorial MLD is visible. The mean reduced bias of the global MLD is given for each season by the average of the
Differences between the mixed layer depth given from the new Lf and control in the coupled simulations during summer and winter in both hemispheres, as in Fig. 5. Note the different color bars to emphasize regional variability. Positive and negative values relate to MLD deepening and shoaling, respectively. The mean reduced bias is given for each season by the average of the
Citation: Journal of Physical Oceanography 53, 1; 10.1175/JPO-D-21-0297.1
As in Fig. 7, but for the forced simulations.
Citation: Journal of Physical Oceanography 53, 1; 10.1175/JPO-D-21-0297.1
5. Summary and discussion
Submesoscales and boundary layer turbulence are instrumental in modulating the transfer of heat, momentum, carbon, and other properties between the atmosphere and ocean interior. Accurate representation of these processes in models is crucial, yet they tend to be on scales smaller than the grid used, even at the highest possible resolution. The current MLE parameterization represents the restratification process of adjusting submesoscale fronts, but it has been shown to be too simplistic in circumstances where the frontal width effects are impactful.
Here we propose a new scaling law that relates frontal width with boundary layer turbulence by building on the TTW balance. The new frontal width scaling utilizes variables from boundary layer turbulence schemes
The new scaling also depends on the local Rossby, Ekman, and turbulent Richardson numbers. Several physical assumptions were made to reduce the number of dependent variables, of which the Rossby and Ekman numbers are assumed to be 1, as expected for the submesoscale. The Richardson number is set to 0.25, representing the arrest by horizontal shear instability.
We test this new scaling over a variety of turbulent processes resulting from winds, waves, and convection that lead to arrested submesoscale fronts and filaments in LES. The predicted frontal width from the new scaling is found be in the same order of the measured frontal width for all cases. Additionally, boundary layer data from the GOTM over four active submesoscale regions was used to estimate the possible range for the new frontal width scaling, which was found to be qualitatively consistent with the submesoscale range of O(1) km. Global estimates of the new scaling were found using a unique configuration of the submesoscale-permitting MITgcm-LLC4320 in tandem with GOTM from Dong et al. (2020), which displayed fronts as sharp as 10 m and up to roughly 50 km near the equator. An exciting future prospect is continuing to evaluate the scaling in observed fronts, perhaps even from satellite datasets (e.g., Ullman et al. 2007; Rascle et al. 2020).
The new scaling for frontal width is implemented in the MLE parameterization and tested in forced and coupled CESM2.1.3-POP simulations, where climate sensitivity was primarily estimated through the impact on the mixed layer depth. Since the new parameterization depends on surface forcing, there is merit in both types of simulations: the coupled simulations have active feedback with the atmosphere, which come back through
The MLE streamfunction is stronger for sharper fronts, that is, for smaller Lf. The frontal width range found by the GOTM and MITgcm-LLC4320 data demonstrates that the new scaling is able to arrive at smaller Lf than predicted by the previous scaling, consistent with the amazingly sharp fronts observed. This will enhance the restratification effect, which is especially important in regions where there is still a bias toward deeper mixed layer depths (e.g., near the equator in summertime).
The LES analysis has demonstrated how important forcing effects such as Langmuir turbulence corrections and penetrating solar corrections are, so adjustments already made in boundary layer schemes, such as a wave-driven turbulence enhancement factor, are readily adapted. This enhancement factor is also available in turbulent mixing schemes in CESM, and future work will include incorporating the effects of Langmuir turbulence in the MLE parameterization. It can be expected that the contribution of Langmuir turbulence will be most significant in the Southern Ocean, where waves tend to be large (Young 1994; Belcher et al. 2012).
Dong et al. (2020, 2021) estimate the scale of symmetric and mixed layer instabilities globally, yet on a local scale it is presently unclear if one dominates over the other and if one is more likely to occur if both conditions are favorable. The Richardson number parameter in the new frontal width scaling is representative of the arrest mechanism. If the front is arrested before reaching horizontal shear instability, the Richardson number could be on order 1 or larger, which may result in frontal widths much larger than currently predicted. Furthermore, if symmetric instability is an important arrest mechanism, it is expected that the MLE parameterization commute with the symmetric instability parameterization and mixing properties given by Bachman et al. (2017).
The new scaling for frontal width was tested using LES of arrested filament frontogenesis, which was treated as two separate fronts for the purposes here. However, the dynamics of filaments have been shown to be different in several aspects, especially in the presence of winds and waves (e.g., Suzuki et al. 2016; Sullivan and McWilliams 2018, 2019). The current MLE parameterization represents the restratifying process of a slumping front, yet MLE may also form along submesoscale filaments if there is sufficient available potential energy stored in the vertical isopycnals. How significant this is, how likely this occurs in nature, and whether the MLE parameterization needs to be modified to include these effects is left for future work.
Frontogenesis is ubiquitous at the submesoscale in the surface layer, whether because of direct action on mesoscale surface horizontal buoyancy gradients or as a secondary frontogenesis in mixed layer eddies that originate from surface-layer baroclinic instability, and whether due to the ambient strain rate or TTW. The end state of frontogenesis is frontal arrest at a finite scale, usually with large Reynolds number, followed by frontal decay. In this paper we propose scaling estimates for the horizontal scale of frontal arrest Lf and the eddy-induced streamfunction Ψ that are expressed entirely in terms of coarse-grid quantities available in climate models. The premises of the proposal are twofold: 1) arrest involves frontal instabilities of various types that limit the horizontal and vertical shear of the front to O(1) bulk values of Ro and Ri, and 2) boundary layer-turbulence on scales smaller than the front are usually involved in the arrest, leading to an O(1) value of Ek (as in the classical Ekman model with
This study intends to demonstrate plausibility, and a proper parameter analysis needs to be performed in order to evaluate the parameterization with simulations that are specifically designed for it. As shown in the supplemental material, fronts can also be arrested by other type of turbulence not captured in the TTW framework (e.g., smaller MLEs without surface forcing). Furthermore, the stress imposed by wind effects can either lead to sharpening or spreading of the front depending on the wind direction, which is not accounted for in this framework. A more comprehensive parameter search is needed to evaluate and compare our results with other scalings for frontal widths under different regimes (e.g., Mahadevan et al. 2010; Wenegrat et al. 2018; Crowe and Taylor 2020). Furthermore, turbulent dissipation rate and frontal width are often a measured observational product. Comparing a wide set of fronts of different sizes, dissipation and surface forcing may help support or counter this scaling.
From CESM3 on, CESM will primarily include the Modular Ocean Model (MOM6). This model includes the ePBL mixing scheme, which allows the use of the precise values of
In the limit that f → 0 the new scaling predicts that Lf becomes increasingly large. This limit is also analogous to the case of a truly nonrotating front (e.g., dam break, Özgökmen et al. 2007), where there is no steady frontal width, as geostrophic balance and thermal wind balance do not constrain the result to remain finite. Thus, Δs/Lf → 0 as the equator is approached, which means that the parameterization shuts itself off as the MLE scale becomes resolved (e.g., Dong et al. 2020). Furthermore, climate models often refine their grid as to resolve tropical instability waves, which resemble MLI in energy source and vertical location (e.g., Danabasoglu et al. 2012), and thus an MLE parameterization active together with resolved mixed layer instabilities would lead to erroneous and nonphysical estimates.
Estimating changes in the mixed layer depth due to the new MLE parameterization gives the most direct impact on climate sensitivity (e.g., Tsujino et al. 2020; Hall and Fox-Kemper 2021, manuscript submitted to Geophys. Res. Lett.). However, there are several other metrics of significant broader climatic impacts that should be evaluated. These include air–sea fluxes and gas exchanges such as CFCs for the impact on the biogeochemical cycle, surface and subsurface tracers such as temperature and salinity, “ideal age” for the effect on ventilation pathways of different water masses below the mixed layer, and on the global scale through the meridional overturning circulation, deep water formation, and average global temperatures. A more comprehensive study is in planning for a more complete application, where model bias and sensitivity are compared with observations in a multimodel multiresolution experiment that may include coupled and JRA55-do forced simulations from CESM-MOM6, CESM-POP, and GFDL-MOM6. Efforts to implement new versions of the MLE parameterization in MOM6 and other models are ongoing.
Acknowledgments.
ASB, PSH, and BFK’s research was made possible in part by a grant from The Gulf of Mexico Research Initiative, and in part by ONR N00014-17-1-2963. LJ, BFK, and PSH were also supported by ONR N00014-17-1-2393. Conversations with Qing Li, Stephen Griffies, and Alistair Adcroft are happily acknowledged as improving this manuscript.
Data availability statement.
Data and code used in the production of this work are available at https://doi.org/10.26300/sf8w-0045.
REFERENCES
Bachman, S., and B. Fox-Kemper, 2013: Eddy parameterization challenge suite I: Eady spindown. Ocean Modell., 64, 12–28, https://doi.org/10.1016/j.ocemod.2012.12.003.
Bachman, S., B. Fox-Kemper, and F. O. Bryan, 2015: A tracer-based inversion method for diagnosing eddy-induced diffusivity and advection. Ocean Modell., 86, 1–14, https://doi.org/10.1016/j.ocemod.2014.11.006.
Bachman, S., B. Fox-Kemper, J. R. Taylor, and L. N. Thomas, 2017: Parameterization of frontal symmetric instabilities. I: Theory for resolved fronts. Ocean Modell., 109, 72–95, https://doi.org/10.1016/j.ocemod.2016.12.003.
Belcher, S. E., and Coauthors, 2012: A global perspective on Langmuir turbulence in the ocean surface boundary layer. Geophys. Res. Lett., 39, L18605, https://doi.org/10.1029/2012GL052932.
Boccaletti, G., R. Ferrari, and B. Fox-Kemper, 2007: Mixed layer instabilities and restratification. J. Phys. Oceanogr., 37, 2228–2250, https://doi.org/10.1175/JPO3101.1.
Bodner, A. S., and B. Fox-Kemper, 2020: A breakdown in potential vorticity estimation delineates the submesoscale-to-turbulence boundary in large eddy simulations. J. Adv. Model. Earth Syst., 12, e2020MS002049, https://doi.org/10.1029/2020MS002049.
Bodner, A. S., B. Fox-Kemper, L. P. Van Roekel, J. C. McWilliams, and P. P. Sullivan, 2020: A perturbation approach to understanding the effects of turbulence on frontogenesis. J. Fluid Mech., 883, A25, https://doi.org/10.1017/jfm.2019.804.
Callies, J., and R. Ferrari, 2018a: Baroclinic instability in the presence of convection. J. Phys. Oceanogr., 48, 45–60, https://doi.org/10.1175/JPO-D-17-0028.1.
Callies, J., and R. Ferrari, 2018b: Note on the rate of restratification in the baroclinic spindown of fronts. J. Phys. Oceanogr., 48, 1543–1553, https://doi.org/10.1175/JPO-D-17-0175.1.
Calvert, D., G. Nurser, M. J. Bell, and B. Fox-Kemper, 2020: The impact of a parameterisation of submesoscale mixed layer eddies on mixed layer depths in the NEMO ocean model. Ocean Modell., 154, 101678, https://doi.org/10.1016/j.ocemod.2020.101678.
Capet, X., J. C. McWilliams, M. J. Molemaker, and A. F. Shchepetkin, 2008: Mesoscale to submesoscale transition in the California Current System. Part I: Flow structure, eddy flux, and observational tests. J. Phys. Oceanogr., 38, 29–43, https://doi.org/10.1175/2007JPO3671.1.
Cronin, M. F., and W. S. Kessler, 2009: Near-surface shear flow in the tropical Pacific cold tongue front. J. Phys. Oceanogr., 39, 1200–1215, https://doi.org/10.1175/2008JPO4064.1.
Crowe, M., and J. R. Taylor, 2020: The effects of surface wind stress and buoyancy flux on the evolution of a front in a turbulent thermal wind balance. Fluids, 5, 87, https://doi.org/10.3390/fluids5020087.
Dale, A. C., J. A. Barth, M. D. Levine, and J. A. Austin, 2008: Observations of mixed layer restratification by onshore surface transport following wind reversal in a coastal upwelling region. J. Geophys. Res., 113, C01010, https://doi.org/10.1029/2007JC004128.
Danabasoglu, G., S. C. Bates, B. P. Briegleb, S. R. Jayne, M. Jochum, W. G. Large, S. Peacock, and S. G. Yeager, 2012: The CCSM4 ocean component. J. Climate, 25, 1361–1389, https://doi.org/10.1175/JCLI-D-11-00091.1.
Dauhajre, D. P., and J. C. McWilliams, 2018: Diurnal evolution of submesoscale front and filament circulations. J. Phys. Oceanogr., 48, 2343–2361, https://doi.org/10.1175/JPO-D-18-0143.1.
de Boyer Montégut, C., G. Madec, A. S. Fischer, A. Lazar, and D. Iudicone, 2004: Mixed layer depth over the global ocean: An examination of profile data and a profile-based climatology. J. Geophys. Res., 109, C12003, https://doi.org/10.1029/2004JC002378.
Dong, J., B. Fox-Kemper, H. Zhang, and C. Dong, 2020: The scale of submesoscale baroclinic instability globally. J. Phys. Oceanogr., 50, 2649–2667, https://doi.org/10.1175/JPO-D-20-0043.1.
Dong, J., B. Fox-Kemper, H. Zhang, and C. Dong, 2021: The scale and activity of symmetric instability estimated from a global submesoscale-permitting ocean model. J. Phys. Oceanogr., 51, 1655–1670, https://doi.org/10.1175/JPO-D-20-0159.1.
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, 2011: Parameterization of mixed layer eddies. III: Implementation and impact in global ocean climate simulations. Ocean Modell., 39, 61–78, https://doi.org/10.1016/j.ocemod.2010.09.002.
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.
Fox-Kemper, B., L. Johnson, and F. Qiao, 2022: Ocean near-surface layers. Ocean Mixing, M. Meredith and A. Naveira Garabato, Eds., Elsevier, 65–94.
Garrett, C. J. R., and J. W. Loder, 1981: Dynamical aspects of shallow sea fronts. Philos. Trans. Roy. Soc., A302, 563–581, https://doi.org/10.1098/rsta.1981.0183.
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. C. McWilliams, and C. Snyder, 1994: Scaling analysis of curved fronts. Validity of the balance equations and semigeostrophy. J. Atmos. Sci., 51, 160–163, https://doi.org/10.1175/1520-0469(1994)051<0160:SAOCFV>2.0.CO;2.
Grant, A. L. M., and S. E. Belcher, 2009: Characteristics of Langmuir turbulence in the ocean mixed layer. J. Phys. Oceanogr., 39, 1871–1887, https://doi.org/10.1175/2009JPO4119.1.
Griffies, S. M., 1998: The Gent–McWilliams skew flux. J. Phys. Oceanogr., 28, 831–841, https://doi.org/10.1175/1520-0485(1998)028<0831:TGMSF>2.0.CO;2.
Gula, J., M. J. Molemaker, and J. C. McWilliams, 2014: Submesoscale cold filaments in the gulf stream. J. Phys. Oceanogr., 44, 2617–2643, https://doi.org/10.1175/JPO-D-14-0029.1.
Gula, J., M. J. Molemaker, and J. C. McWilliams, 2016: Topographic generation of submesoscale centrifugal instability and energy dissipation. Nat. Commun., 7, 12811, https://doi.org/10.1038/ncomms12811.
Haine, T. W. N., and J. Marshall, 1998: Gravitational, symmetric, and baroclinic instability of the ocean mixed layer. J. Phys. Oceanogr., 28, 634–658, https://doi.org/10.1175/1520-0485(1998)028<0634:GSABIO>2.0.CO;2.
Hamlington, P. E., L. P. Van Roekel, B. Fox-Kemper, K. Julien, and G. P. Chini, 2014: Langmuir–submesoscale interactions: Descriptive analysis of multiscale frontal spindown simulations. J. Phys. Oceanogr., 44, 2249–2272, https://doi.org/10.1175/JPO-D-13-0139.1.
Haney, S., B. Fox-Kemper, K. Julien, and A. Webb, 2015: Symmetric and geostrophic instabilities in the wave-forced ocean mixed layer. J. Phys. Oceanogr., 45, 3033–3056, https://doi.org/10.1175/JPO-D-15-0044.1.
Hosegood, P., M. C. Gregg, and M. H. Alford, 2006: Sub-mesoscale lateral density structure in the oceanic surface mixed layer. Geophys. Res. Lett., 33, L22604, https://doi.org/10.1029/2006GL026797.
Hoskins, B. J., 1974: The role of potential vorticity in symmetric stability and instability. Quart. J. Roy. Meteor. Soc., 100, 480–482, https://doi.org/10.1002/qj.49710042520.
Hoskins, B. J., and F. P. Bretherton, 1972: Atmospheric frontogenesis models: Mathematical formulation and solution. J. Atmos. Sci., 29, 11–37, https://doi.org/10.1175/1520-0469(1972)029<0011:AFMMFA>2.0.CO;2.
Johnson, L., C. M. Lee, E. A. D’Asaro, L. Thomas, and A. Shcherbina, 2020: Restratification at a California current upwelling front. Part I: Observations. J. Phys. Oceanogr., 50, 1455–1472, https://doi.org/10.1175/JPO-D-19-0203.1.
Kraus, E. B., and J. S. Turner, 1967: A one-dimensional model of the seasonal thermocline II. The general theory and its consequences. Tellus, 19, 98–106, https://doi.org/10.3402/tellusa.v19i1.9753.
Large, W. G., and S. G. Yeager, 2009: The global climatology of an interannually varying air–sea flux data set. Climate Dyn., 33, 341–364, https://doi.org/10.1007/s00382-008-0441-3.
Large, W. G., J. C. McWilliams, and S. C. Doney, 1994: Oceanic vertical mixing: A review and a model with a nonlocal boundary layer parameterization. Rev. Geophys., 32, 363–403, https://doi.org/10.1029/94RG01872.
Levy, M., G. Danabasoglu, S. Griffies, T. Ringler, A. Adcroft, R. Hallberg, D. Jacobsen, and W. Large, 2014: Community ocean vertical mixing (CVMix) parameterizations. Ocean Model Working Group Meeting, Boulder, CO, UCAR, https://www.cesm.ucar.edu/events/wg-meetings/2014/presentations/omwg/levy.pdf.
Lévy, M., P. J. S. Franks, and K. S. Smith, 2018: The role of submesoscale currents in structuring marine ecosystems. Nat. Commun., 9, 4758, https://doi.org/10.1038/s41467-018-07059-3.
Li, Q., and B. Fox-Kemper, 2017: Assessing the effects of Langmuir turbulence on the entrainment buoyancy flux in the ocean surface boundary layer. J. Phys. Oceanogr., 47, 2863–2886, https://doi.org/10.1175/JPO-D-17-0085.1.
Li, Q., and Coauthors, 2019: Comparing ocean surface boundary vertical mixing schemes including Langmuir turbulence. J. Adv. Model. Earth Syst., 11, 3545–3592, https://doi.org/10.1029/2019MS001810.
Li, Q., J. Bruggeman, H. Burchard, K. Klingbeil, L. Umlauf, and K. Bolding, 2021: Integrating CVMIX into GOTM (v6. 0): A consistent framework for testing, comparing, and applying ocean mixing schemes. Geosci. Model Dev., 14, 4261–4282, https://doi.org/10.5194/gmd-14-4261-2021.
Lucas, A. J., and Coauthors, 2016: Adrift upon a salinity-stratified sea: A view of upper-ocean processes in the bay of bengal during the southwest monsoon. Oceanography, 29, 134–145, https://doi.org/10.5670/oceanog.2016.46.
Mahadevan, A., 2016: The impact of submesoscale physics on primary productivity of plankton. Annu. Rev. Mar. Sci., 8, 161–184, https://doi.org/10.1146/annurev-marine-010814-015912.
Mahadevan, A., A. Tandon, and R. Ferrari, 2010: Rapid changes in mixed layer stratification driven by submesoscale instabilities and winds. J. Geophys. Res., 115, C03017, https://doi.org/10.1029/2008JC005203.
McWilliams, J. C., 2016: Submesoscale currents in the ocean. Proc. Roy. Soc., 472A, 20160117, https://doi.org/10.1098/rspa.2016.0117.
McWilliams, J. C., 2017: Submesoscale surface fronts and filaments: Secondary circulation, buoyancy flux, and frontogenesis. J. Fluid Mech., 823, 391–432, https://doi.org/10.1017/jfm.2017.294.
McWilliams, J. C., 2021: Oceanic frontogenesis. Annu. Rev. Mar. Sci., 13, 227–253, https://doi.org/10.1146/annurev-marine-032320-120725.
McWilliams, J. C., and J. M. Molemaker, 2011: Baroclinic frontal arrest: A sequel to unstable frontogenesis. J. Phys. Oceanogr., 41, 601–619, https://doi.org/10.1175/2010JPO4493.1.
McWilliams, J. C., P. P. Sullivan, and C.-H. Moeng, 1997: Langmuir turbulence in the ocean. J. Fluid Mech., 334, 1–30, https://doi.org/10.1017/S0022112096004375.
McWilliams, J. C., M. J. Molemaker, and E. I. Olafsdottir, 2009: Linear fluctuation growth during frontogenesis. J. Phys. Oceanogr., 39, 3111–3129, https://doi.org/10.1175/2009JPO4186.1.
McWilliams, J. C., J. Gula, J. M. Molemaker, L. Renault, and A. F. Shchepetkin, 2015: Filament frontogenesis by boundary layer turbulence. J. Phys. Oceanogr., 45, 1988–2005, https://doi.org/10.1175/JPO-D-14-0211.1.
Mironov, D., A. Terzhevik, G. Kirillin, T. Jonas, J. Malm, and D. Farmer, 2002: Radiatively driven convection in ice-covered lakes: Observations, scaling, and a mixed layer model. J. Geophys. Res., 107, 3032, https://doi.org/10.1029/2001JC000892.
Nagai, T., A. Tandon, and D. L. Rudnick, 2006: Two-dimensional ageostrophic secondary circulation at ocean fronts due to vertical mixing and large-scale deformation. J. Geophys. Res., 111, C09038, https://doi.org/10.1029/2005JC002964.
Nurser, A. J. G., and J. W. Zhang, 2000: Eddy-induced mixed layer shallowing and mixed layer/thermocline exchange. J. Geophys. Res., 105, 21 851–21 868, https://doi.org/10.1029/2000JC900018.
Olita, A., and Coauthors, 2017: Frontal dynamics boost primary production in the summer stratified mediterranean sea. Ocean Dyn., 67, 767–782, https://doi.org/10.1007/s10236-017-1058-z.
Özgökmen, T. M., T. Iliescu, P. F. Fischer, A. Srinivasan, and J. Duan, 2007: Large eddy simulation of stratified mixing in two-dimensional dam-break problem in a rectangular enclosed domain. Ocean Modell., 16, 106–140, https://doi.org/10.1016/j.ocemod.2006.08.006.
Pallàs-Sanz, E., T. M. S. Johnston, and D. L. Rudnick, 2010: Frontal dynamics in a California current system shallow front: 1. Frontal processes and tracer structure. J. Geophys. Res., 115, C12067, https://doi.org/10.1029/2009JC006032.
Price, J. F., R. A. Weller, and R. Pinkel, 1986: Diurnal cycling: Observations and models of the upper ocean response to diurnal heating, cooling, and wind mixing. J. Geophys. Res., 91, 8411–8427, https://doi.org/10.1029/JC091iC07p08411.
Ramachandran, S., and Coauthors, 2018: Submesoscale processes at shallow salinity fronts in the Bay of Bengal: Observations during the winter monsoon. J. Phys. Oceanogr., 48, 479–509, https://doi.org/10.1175/JPO-D-16-0283.1.
Rascle, N., and Coauthors, 2020: Monitoring intense oceanic fronts using sea surface roughness: Satellite, airplane, and in situ comparison. J. Geophys. Res. Oceans, 125, e2019JC015704, https://doi.org/10.1029/2019JC015704.
Reichl, B. G., and R. Hallberg, 2018: A simplified energetics based planetary boundary layer (ePBL) approach for ocean climate simulations. Ocean Modell., 132, 112–129, https://doi.org/10.1016/j.ocemod.2018.10.004.
Rocha, C. B., T. K. Chereskin, S. T. Gille, and D. Menemenlis, 2016: Mesoscale to submesoscale wavenumber spectra in Drake Passage. J. Phys. Oceanogr., 46, 601–620, https://doi.org/10.1175/JPO-D-15-0087.1.
Rotunno, R., W. C. Skamarock, and C. Snyder, 1994: An analysis of frontogenesis in numerical simulations of baroclinic waves. J. Atmos. Sci., 51, 3373–3398, https://doi.org/10.1175/1520-0469(1994)051<3373:AAOFIN>2.0.CO;2.
Skyllingstad, E. D., and R. Samelson, 2012: Baroclinic frontal instabilities and turbulent mixing in the surface boundary layer. Part I: Unforced simulations. J. Phys. Oceanogr., 42, 1701–1716, https://doi.org/10.1175/JPO-D-10-05016.1.
Smith, K. M., P. E. Hamlington, and B. Fox-Kemper, 2016: Effects of submesoscale turbulence on ocean tracers. J. Geophys. Res. Oceans, 121, 908–933, https://doi.org/10.1002/2015JC011089.
Su, Z., J. Wang, P. Klein, A. F. Thompson, and D. Menemenlis, 2018: Ocean submesoscales as a key component of the global heat budget. Nat. Commun., 9, 775, https://doi.org/10.1038/s41467-018-02983-w.
Sullivan, P. P., and J. C. McWilliams, 2018: Frontogenesis and frontal arrest of a dense filament in the oceanic surface boundary layer. J. Fluid Mech., 837, 341–380, https://doi.org/10.1017/jfm.2017.833.
Sullivan, P. P., and J. C. McWilliams, 2019: Langmuir turbulence and filament frontogenesis in the oceanic surface boundary layer. J. Fluid Mech., 879, 512–553, https://doi.org/10.1017/jfm.2019.655.
Sullivan, P. P., J. C. McWilliams, and W. K. Melville, 2007: Surface gravity wave effects in the oceanic boundary layer: Large-eddy simulation with vortex force and stochastic breakers. J. Fluid Mech., 593, 405–452, https://doi.org/10.1017/S002211200700897X.
Suzuki, N., B. Fox-Kemper, P. E. Hamlington, and L. P. Van Roekel, 2016: Surface waves affect frontogenesis. J. Geophys. Res. Oceans, 121, 3597–3624, https://doi.org/10.1002/2015JC011563.
Tandon, A., and C. Garrett, 1994: Mixed layer restratification due to a horizontal density gradient. J. Phys. Oceanogr., 24, 1419–1424, https://doi.org/10.1175/1520-0485(1994)024<1419:MLRDTA>2.0.CO;2.
Taylor, J. R., and R. Ferrari, 2011: Ocean fronts trigger high latitude phytoplankton blooms. Geophys. Res. Lett., 38, L23601, https://doi.org/10.1029/2011GL049312.
Teixeira, M. C., and S. E. Belcher, 2002: On the distortion of turbulence by a progressive surface wave. J. Fluid Mech., 458, 229–267, https://doi.org/10.1017/S0022112002007838.
Thomas, L. N., J. R. Taylor, R. Ferrari, and T. M. Joyce, 2013: Symmetric instability in the gulf stream. Deep-Sea Res. II, 91, 96–110, https://doi.org/10.1016/j.dsr2.2013.02.025.
Thompson, A. F., A. Lazar, C. Buckingham, A. C. Naveira Garabato, G. M. Damerell, and K. J. Heywood, 2016: Open-ocean submesoscale motions: A full seasonal cycle of mixed layer instabilities from gliders. J. Phys. Oceanogr., 46, 1285–1307, https://doi.org/10.1175/JPO-D-15-0170.1.
Thompson, L., 2000: Ekman layers and two-dimensional frontogenesis in the upper ocean. J. Geophys. Res., 105, 6437–6451, https://doi.org/10.1029/1999JC900336.
Thorpe, S. A., 2005: The Turbulent Ocean. Cambridge University Press, 496 pp.
Timmermans, M.-L., and P. Winsor, 2013: Scales of horizontal density structure in the Chukchi Sea surface layer. Cont. Shelf Res., 52, 39–45, https://doi.org/10.1016/j.csr.2012.10.015.
Tsujino, H., and Coauthors, 2020: Evaluation of global ocean–sea-ice model simulations based on the experimental protocols of the Ocean Model Intercomparison Project phase 2 (OMIP-2). Geosci. Model Dev., 13, 3643–3708, https://doi.org/10.5194/gmd-13-3643-2020.
Ullman, D. S., P. C. Cornillon, and Z. Shan, 2007: On the characteristics of subtropical fronts in the North Atlantic. J. Geophys. Res., 112, C01010, https://doi.org/10.1029/2006JC003601.
Umlauf, L., and H. Burchard, 2005: Second-order turbulence closure models for geophysical boundary layers. A review of recent work. Cont. Shelf Res., 25, 795–827, https://doi.org/10.1016/j.csr.2004.08.004.
Van Roekel, L. P., P. E. Hamlington, and B. Fox-Kemper, 2012: Multiscale simulations of Langmuir cells and submesoscale eddies using XSEDE resources. Proc. XSEDE12 First Conf. of the Extreme Science and Engineering Discovery Environment, Chicago, IL, ACM, 20, https://doi.org/10.1145/2335755.2335816.
Van Roekel, L. P., and Coauthors, 2018: The KPP boundary layer scheme for the ocean: Revisiting its formulation and benchmarking one-dimensional simulations relative to LES. J. Adv. Model. Earth Syst., 10, 2647–2685, https://doi.org/10.1029/2018MS001336.
Wenegrat, J. O., L. N. Thomas, J. Gula, and J. C. McWilliams, 2018: Effects of the submesoscale on the potential vorticity budget of ocean mode waters. J. Phys. Oceanogr., 48, 2141–2165, https://doi.org/10.1175/JPO-D-17-0219.1.
Young, I. R., 1994: Global ocean wave statistics obtained from satellite observations. Appl. Ocean Res., 16, 235–248, https://doi.org/10.1016/0141-1187(94)90023-X.