1. Introduction
The current generation of climate models shows significant biases in simulating the sea surface temperature and mixed layer depth that are key to the air–sea interactions (Fox-Kemper et al. 2021; Huang et al. 2014). One possible cause of such biases is the incapability of resolving oceanic small-scale processes at O(0.1–100) km due to the coarse model resolution. By generating a prominent restratifying vertical buoyancy flux (VBF), these small-scale processes act to warm the sea surface and shallow the surface mixed layer (Boccaletti et al. 2007; Callies and Ferrari 2018; Fox-Kemper et al. 2011; Jing et al. 2020; Su et al. 2018). A reliable parameterization of their restratification effect is thus crucial to improve the fidelity of climate simulations without increasing the model resolution and computational burden.
The mixed layer instability (MLI), frontogenesis, symmetric instability (SI), and turbulent thermal wind (TTW) balance are considered the major mechanisms for the VBF of small-scale processes in the surface mixed layer (Mahadevan and Tandon 2006; McWilliams 2016; Yang et al. 2021a). The MLI is a type of baroclinic instability occurring in the surface mixed layer (Boccaletti et al. 2007; Fox-Kemper et al. 2008), converting the available potential energy stored in the mixed layer fronts to kinetic energy. The frontogenesis occurs at an intensifying front caused by background confluent flows, inducing an ageostrophic secondary circulation (ASC) with upwelling and downwelling on the lighter and denser side of the front, respectively (Barkan et al. 2019; Hoskins 1982; McWilliams 2021). The SI is triggered if the Ertel potential vorticity (PV) has an opposite sign to the planetary vorticity and tends to occur in the surface frontal regions when the PV is extracted from the ocean due to the surface buoyancy loss (D’Asaro et al. 2011; Jing et al. 2021; Thomas 2005; Thomas et al. 2013). In readjusting to a symmetrically neutral state, the SI restratifies the surface mixed layer with the magnitude of its VBF related to the surface buoyancy loss (Bachman et al. 2017; Taylor and Ferrari 2010). The TTW balance is a linear momentum balance between Coriolis force, horizontal pressure gradient, and turbulent vertical mixing (Gula et al. 2014; Wenegrat and McPhaden 2016; Yang et al. 2021a). The viscous effect destroys the vertical shear of the front, inducing a restoring ASC that has a similar pattern to that of the frontogenesis (Gula et al. 2014; McWilliams 2016; McWilliams et al. 2015).
Parameterizations of the VBF induced by the MLI, frontogenesis, and SI have been proposed (Bachman et al. 2017; Fox-Kemper et al. 2011; Zhang et al. 2023). Among them, the parameterization of the MLI (Bodner et al. 2023; Fox-Kemper et al. 2011) is relatively more mature and has been widely applied in climate simulations (Calvert et al. 2020; Chang et al. 2020; Fox-Kemper et al. 2011; Griffies et al. 2015; G. Xu et al. 2022). This parameterization is scale aware and formulated as an overturning streamfunction related to the horizontal density gradient in the surface mixed layer scaled by model resolution, the mixed layer depth, and the Coriolis frequency (Bodner et al. 2023; Fox-Kemper et al. 2011, 2008; Fox-Kemper and Ferrari 2008). However, the parameterization of the VBF of the TTW balance is still lacking. Yet, a numerical study in the winter Kuroshio Extension (Yang et al. 2021a) suggests that the VBF of the TTW balance is comparable to the sum of those induced by the MLI and frontogenesis. In this study, we aim to propose a scale-aware parameterization for the VBF of the TTW balance and test its performance in a set of ocean simulations in the winter Kuroshio Extension with horizontal resolutions varying from 1 to 27 km. The VBF of the TTW balance is well resolved in the 1-km simulation, where the 27 km is close to the highest resolution affordable for long-term global climate simulations nowadays (Haarsma et al. 2016; Hewitt et al. 2020). The manuscript is organized as follows. The configuration of numerical models is described in section 2. Section 3 proposes the parameterization and evaluates its performance. Discussion on the parameterization is given in section 4 followed by conclusions in section 5.
2. Numerical experiments
To evaluate the performance of the parameterization for the VBF of the TTW balance, the Regional Ocean Modeling System (ROMS) (Haidvogel et al. 2000; Shchepetkin and McWilliams 2005) is used to configure a series of offline-nested simulations over the North Pacific focused on the Kuroshio Extension. The lowest-resolution simulation, named NP27, has a modeling domain over the North Pacific (99°–270°E, 3.6°–66°N; Fig. 1a), with a horizontal resolution of ∼27 km and 50 terrain-following vertical levels. The initial state and boundary conditions for the NP27 are obtained from the Simple Ocean Data Assimilation (SODA; Carton et al. 2018). The simulation is initialized on 1 January 1998 and integrated to 31 March 2004, outputting 3-hourly snapshots of temperature, salinity, three-dimensional velocity, and turbulent vertical viscosity as well as daily averaged diagnostic terms in momentum and tracer equations.
Snapshots of the normalized relative vertical vorticity ς/f in the (a) NP27, (b) KE9, (c) KE3, and (d) KE1. The model domain of the KE9, KE3, and KE1 is enclosed by black lines in (a), and the region used for validating the parameterization is enclosed by black lines in (d).
Citation: Journal of Physical Oceanography 54, 5; 10.1175/JPO-D-23-0169.1
The finer simulations in the hierarchy at ∼9-, ∼3-, and ∼1-km resolution are configured over the Kuroshio Extension (named KE9, KE3, and KE1 hereinafter). The domains cover 142.5°–173°E, 27°–46°N for the KE9, 147°–170°E, 29°–44.5°N for the KE3, and 148°–168°E, 31°–43.5°N for the KE1, shrinking with the increasing resolution (Figs. 1b–d). The vertical levels of the KE9 are the same as those of the NP27, while the KE3 and KE1 have 65 vertical levels with ∼5-m grid size in the upper 100 m. The initial states and the one-way nested open boundary conditions for each KE simulation are obtained from the 3-hourly snapshots of its immediate parent simulation. After the spinup of their immediate parent simulations, the KE9, KE3, and KE1 are initialized on 1 January, 1 July, and 1 October 2003, respectively. All these three simulations are run until 31 March 2004.
In all the simulations, the horizontal and vertical advections for momentum are discretized using a third-order upwind scheme and fourth-order central difference scheme, respectively. The multidimensional positive definite advection transport algorithm (MPDATA; Margolin and Smolarkiewicz 1998) is used for the horizontal and vertical advection of tracers. We use the K-profile parameterization (KPP) turbulent mixing closure scheme for the vertical mixing of momentum and tracers (Large et al. 1994) and a biharmonic horizontal Smagorinsky-like mixing scheme for momentum (Griffies and Hallberg 2000; Smagorinsky 1963). A biharmonic horizontal mixing scheme of tracers is used in the NP27 with diffusivity set as −1 × 1010 m4 s−1, whereas no horizontal mixing is used for tracers in the rest simulations. In all the simulations, the atmospheric forcing is calculated from a bulk formula based on the 6-hourly Climate Forecast System Reanalysis (CFSR; Saha et al. 2010), and the sea surface salinity is restored to that in the SODA with a nudging time scale of 10 days. In addition to the 3-hourly snapshots, all the KE simulations output the 3-hourly averaged diagnostic terms in momentum and tracer equations. The above model configurations are summarized in Table 1.
A summary of numerical configurations of the NP27, KE9, KE3, and KE1.
Zhang et al. (2023) reported that the VBF of the TTW balance has strong seasonality with its value in summer one order of magnitude smaller than that in winter. Therefore, we focus on the performance of the parameterization in winter, based on the model output from 1 December 2003 to 31 March 2004 in an inner domain of the KE1, i.e., 150°–166°E, 33°–42°N (Fig. 1d).
3. A scale-aware parameterization of the VBF induced by the TTW balance
a. VBF induced by the TTW balance
Equation (1) models the ASCs driven by both τw and the geostrophic stress τg = Km∂ug/∂z (Cronin and Kessler 2009; F. Xu et al. 2022; Yang et al. 2021b). The former can generate a positive VBF via the current feedback (Dawe and Thompson 2006; Duhaut and Straub 2006), but its magnitude is generally much smaller than that induced by τg at least in winter in the Kuroshio Extension region (F. Xu et al. 2022; Yang et al. 2021b). In the following parameterization, we focus only on the VBF induced by τg by setting τw = 0.
Vertical profiles of model-resolved (a) turbulent vertical viscosity
Citation: Journal of Physical Oceanography 54, 5; 10.1175/JPO-D-23-0169.1
The magnitude of
b. Scaling of the horizontal density gradient
The above analysis suggests that a scale-aware parameterization of
The squared horizontal density gradient averaged within the mixed layer
Citation: Journal of Physical Oceanography 54, 5; 10.1175/JPO-D-23-0169.1
In the winter Kuroshio Extension,
(a) Probability density function (PDF) of the mixed layer deformation radius
Citation: Journal of Physical Oceanography 54, 5; 10.1175/JPO-D-23-0169.1
A list of variables averaged over 150°–166°E, 33°–42°N from 1 Dec 2003 to 31 Mar 2004 in the (a) NP27, KE9, KE3, and KE1 and (b) NP27-P, KE9-P, KE3-P, and KE1-P. The peaking value of
c. Implementation of the parameterization in a numerical model
To implement the parameterization equation [Eq. (9)] in a numerical model like ROMS, we derive
d. Performance of the parameterization in numerical simulations with different resolutions
1) Parameterized VBF under the TTW balance
Equations (10) and (11) are implemented online in the KE1, KE3, KE9, and NP27, costing about 2.7% of the total computational time with the potential for further optimizations. To distinguish numerical simulations with and without the parameterizations, we add a suffix “P” to the former (e.g., NP27-P vs NP27). In the numerical simulations without the parameterization, the difference of
Vertical profiles of the (a) model-resolved VBF under the TTW balance
Citation: Journal of Physical Oceanography 54, 5; 10.1175/JPO-D-23-0169.1
The
2) Restratification effect of the parameterization
The primary role of
(a) Temporal variations of the surface buoyancy loss (black) and wind stress (blue) averaged over 150°–166°E, 33°–42°N. Temporal and vertical variations of (b) horizontally averaged squared buoyancy frequency
Citation: Journal of Physical Oceanography 54, 5; 10.1175/JPO-D-23-0169.1
3) Interaction between resolved and parameterized VBFs under the TTW balance
There is a significant interaction between
Vertical profiles of
Citation: Journal of Physical Oceanography 54, 5; 10.1175/JPO-D-23-0169.1
4. Discussion
a. The overestimation of the parameterization in coarse-resolution simulation
The first term on the right-hand side of Eq. (12) has similar values between the NP27-P and KE1-P (Fig. 8b) because both 〈(∇Hb)2〉 and
Vertical profiles of (a)
Citation: Journal of Physical Oceanography 54, 5; 10.1175/JPO-D-23-0169.1
The above analysis indicates that the scaling of
b. Interactions between the parameterized VBF induced by the TTW balance and MLI instability
By decomposing the VBF induced by different submesoscale dynamical processes (including the MLI and TTW balance), Yang et al. (2021a) find that the MLI- and TTW-induced VBFs coexist and have comparable magnitudes in the winter Kuroshio Extension. However, the MLI- and TTW-induced VBFs are not independent and are likely to interact with each other and so are their parameterizations. These two parameterizations are likely to exert negative feedback on each other, as both the parameterizations act to restratify the surface mixed layer. On the one hand, including the parameterization of TTW-induced VBF shoals the surface mixed layer, diminishing the parameterized MLI-induced VBF as it is proportional to the square of surface mixed layer depth (Fox-Kemper et al. 2008, 2011). On the other hand, including the parameterization of MLI-induced VBF enhances the stratification, reducing the turbulent vertical viscosity and parameterized TTW-induced VBF.
c. Effect of the turbulent vertical mixing schemes on the parameterization
In this study, the parameterization is implemented in simulations with the turbulent vertical mixing parameterized by the KPP scheme. Technically, it can also be implemented in simulations with a generic length scale (GLS) closure scheme, as long as the Reynolds stress can be expressed in the form of Km ∂u/∂z, where Km varies smoothly with depth.
For example, the Mellor–Yamada level 2.5 (MY25) scheme is a widely used GLS scheme in ocean and climate models (Mellor and Yamada 1974). We conduct another 1- and 9-km resolution simulation using the MY25 scheme with the TTW parameterization implemented (KE1-MY25-P and KE9-MY25-P). The KE1-MY25-P and KE9-MY25-P share the same configurations with the KE1-P and KE9-P, respectively, except for the turbulent vertical mixing scheme. As shown in Figs. 9a and 9c, our parameterization works qualitatively well for both the KPP and MY25 schemes.
(a) Vertical profiles of
Citation: Journal of Physical Oceanography 54, 5; 10.1175/JPO-D-23-0169.1
Quantitatively, the performance of our parameterization depends on the turbulent vertical mixing scheme. The magnitude of
d. Width of mixed layer fronts
In this study, the width of mixed layer fronts is estimated as the mixed layer deformation radius following Fox-Kemper et al. (2011). Recently, Bodner et al. (2023) propose another formula to estimate the width of mixed layer fronts, by taking into consideration the arresting effect of turbulent flux on the front width under the TTW balance. Application of Bodner et al.’s (2023) formula to our simulations suggests that the mixed layer fronts can be as sharp as O(100) m, an order of magnitude smaller than the mixed layer deformation radius in the winter Kuroshio Extension. Bodner et al.’s (2023) formula implies that
5. Conclusions
In this study, we propose a scale-aware parameterization for the VBF induced by the TTW balance
-
The parameterization depends on the Coriolis parameter, model-resolved turbulent vertical viscosity, and horizontal density gradient. It employs a scaling relationship as a function of horizontal model resolution and mixed layer deformation radius to adjust for the effect of horizontal model resolution on the simulated width of mixed layer fronts. In particular, the parameterization is automatically inactivated if the mixed layer fronts are resolved in the models.
-
The parameterization shows good skills in reconciling the difference of
across models with different horizontal resolutions. Their peaking values of differ by less than 32% for horizontal resolutions ranging from 27 to 1 km. Implementing the parameterization leads to a better representation of the stratification within the surface mixed layer in coarse-resolution simulations.
There is a moderate but noticeable overestimation of
Acknowledgments.
This research is supported by the National Key Research and Development Program of China (2022YFC3104801), National Natural Science Foundation of China (42206024 and 42176218), Science and Technology Innovation Project of Laoshan Laboratory (LSKJ202202501), and computational resources are supported by Laoshan Laboratory (LSKJ202300302). Supports are also obtained from Taishan Scholar Funds (tsqn201909052). The model simulation and many of the computations were executed at the High Performance Computing Center of Laoshan Laboratory.
Data availability statement.
The simulation data used in this work can be reproduced using the ROMS code available at https://www.myroms.org/projects/src/wiki.
Please note that Eq. (1a) does not appear in the same form as Eq. (12) of Gula et al. (2014). The latter can be derived by taking the vertical derivative of Eq. (1a).
REFERENCES
Bachman, S. D., 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.
Barkan, R., M. J. Molemaker, K. Srinivasan, J. C. McWilliams, and E. A. D’Asaro, 2019: The role of horizontal divergence in submesoscale frontogenesis. J. Phys. Oceanogr., 49, 1593–1618, https://doi.org/10.1175/JPO-D-18-0162.1.
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., B. Fox-Kemper, L. Johnson, L. P. Van Roekel, J. C. McWilliams, P. P. Sullivan, P. S. Hall, and J. Dong, 2023: Modifying the mixed layer eddy parameterization to include frontogenesis arrest by boundary layer turbulence. J. Phys. Oceanogr., 53, 323–339, https://doi.org/10.1175/JPO-D-21-0297.1.
Callies, J., and R. Ferrari, 2018: 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.
Carton, J. A., G. A. Chepurin, and L. Chen, 2018: SODA3: A new ocean climate reanalysis. J. Climate, 31, 6967–6983, https://doi.org/10.1175/JCLI-D-18-0149.1.
Chang, P., and Coauthors, 2020: An unprecedented set of high‐resolution Earth system simulations for understanding multiscale interactions in climate variability and change. J. Adv. Model. Earth Syst., 12, e2020MS002298, https://doi.org/10.1029/2020MS002298.
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.
D’Asaro, E. A., C. Lee, L. Rainville, R. Harcourt, and L. Thomas, 2011: Enhanced turbulence and energy dissipation at ocean fronts. Science, 332, 318–322, https://doi.org/10.1126/science.1201515.
Dawe, J. T., and L. Thompson, 2006: Effect of ocean surface currents on wind stress, heat flux, and wind power input to the ocean. Geophys. Res. Lett., 33, L09604, https://doi.org/10.1029/2006GL025784.
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.
Duhaut, T. H. A., and D. N. Straub, 2006: Wind stress dependence on ocean surface velocity: Implications for mechanical energy input to ocean circulation. J. Phys. Oceanogr., 36, 202–211, https://doi.org/10.1175/JPO2842.1.
Fox-Kemper, B., and R. Ferrari, 2008: Parameterization of mixed layer eddies. Part II: Prognosis and impact. J. Phys. Oceanogr., 38, 1166–1179, https://doi.org/10.1175/2007JPO3788.1.
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, 2021: Ocean, cryosphere and sea level change. Climate Change 2021: The Physical Science Basis, V. Masson-Delmotte et al., Eds., Cambridge University Press, 1211–1362, https://doi.org/10.1017/9781009157896.011.
Griffies, S. M., and R. W. Hallberg, 2000: Biharmonic friction with a Smagorinsky-like viscosity for use in large-scale eddy-permitting ocean models. Mon. Wea. Rev., 128, 2935–2946, https://doi.org/10.1175/1520-0493(2000)128<2935:BFWASL>2.0.CO;2.
Griffies, S. M., and Coauthors, 2015: Impacts on ocean heat from transient mesoscale eddies in a hierarchy of climate models. J. Climate, 28, 952–977, https://doi.org/10.1175/JCLI-D-14-00353.1.
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.
Haarsma, R. J., and Coauthors, 2016: High Resolution Model Intercomparison Project (HighResMIP v1.0) for CMIP6. Geosci. Model Dev., 9, 4185–4208, https://doi.org/10.5194/gmd-9-4185-2016.
Haidvogel, D. B., H. G. Arango, K. Hedstrom, A. Beckmann, P. Malanotte-Rizzoli, and A. F. Shchepetkin, 2000: Model evaluation experiments in the North Atlantic Basin: Simulations in nonlinear terrain-following coordinates. Dyn. Atmos. Oceans, 32, 239–281, https://doi.org/10.1016/S0377-0265(00)00049-X.
Hewitt, H. T., and Coauthors, 2020: Resolving and parameterising the ocean mesoscale in Earth System Models. Curr. Climate Change Rep., 6, 137–152, https://doi.org/10.1007/s40641-020-00164-w.
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.
Huang, C. J., F. Qiao, and D. Dai, 2014: Evaluating CMIP5 simulations of mixed layer depth during summer. J. Geophys. Res. Oceans, 119, 2568–2582, https://doi.org/10.1002/2013JC009535.
Jing, Z., and Coauthors, 2020: Maintenance of mid-latitude oceanic fronts by mesoscale eddies. Sci. Adv., 6, eaba7880, https://doi.org/10.1126/sciadv.aba7880.
Jing, Z., B. Fox-Kemper, H. Cao, R. Zheng, and Y. Du, 2021: Submesoscale fronts and their dynamical processes associated with symmetric instability in the northwest Pacific subtropical ocean. J. Phys. Oceanogr., 51, 83–100, https://doi.org/10.1175/JPO-D-20-0076.1.
Large, W. G., G. Danabasoglu, S. C. Doney, and J. C. McWilliams, 1997: Sensitivity to surface forcing and boundary layer mixing in a global ocean model: Annual-mean climatology. J. Phys. Oceanogr., 27, 2418–2447, https://doi.org/10.1175/1520-0485(1997)027<2418:STSFAB>2.0.CO;2.
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.
Mahadevan, A., 2006: Modeling vertical motion at ocean fronts: Are nonhydrostatic effects relevant at submesoscales? Ocean Modell., 14, 222–240, https://doi.org/10.1016/j.ocemod.2006.05.005.
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.
Margolin, L., and P. K. Smolarkiewicz, 1998: Antidiffusive velocities for multipass donor cell advection. SIAM J. Sci. Comput., 20, 907–929, https://doi.org/10.1137/S106482759324700X.
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., J. Gula, M. J. 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.
Mellor, G. L., and T. Yamada, 1974: A hierarchy of turbulence closure models for planetary boundary layers. J. Atmos. Sci., 31, 1791–1806, https://doi.org/10.1175/1520-0469(1974)031<1791:AHOTCM>2.0.CO;2.
Saha, S., and Coauthors, 2010: The NCEP Climate Forecast System Reanalysis. Bull. Amer. Meteor. Soc., 91, 1015–1058, https://doi.org/10.1175/2010BAMS3001.1.
Shchepetkin, A. F., and J. C. McWilliams, 2005: The Regional Oceanic Modeling System (ROMS): A split-explicit, free-surface, topography-following-coordinate oceanic model. Ocean Modell., 9, 347–404, https://doi.org/10.1016/j.ocemod.2004.08.002.
Smagorinsky, J., 1963: General circulation experiments with the primitive equations: I. The basic experiment. Mon. Wea. Rev., 91, 99–164, https://doi.org/10.1175/1520-0493(1963)091<0099:GCEWTP>2.3.CO;2.
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.
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, 2010: Buoyancy and wind-driven convection at mixed layer density fronts. J. Phys. Oceanogr., 40, 1222–1242, https://doi.org/10.1175/2010JPO4365.1.
Thomas, L. N., 2005: Destruction of potential vorticity by winds. J. Phys. Oceanogr., 35, 2457–2466, https://doi.org/10.1175/JPO2830.1.
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.
Wenegrat, J. O., and M. J. McPhaden, 2016: Wind, waves, and fronts: Frictional effects in a generalized Ekman model. J. Phys. Oceanogr., 46, 371–394, https://doi.org/10.1175/JPO-D-15-0162.1.
Xu, F., Z. Jing, P. Yang, and S. Zhou, 2022: Vertical motions driven by geostrophic stress in surface boundary layer under a finite Ekman number regime. J. Phys. Oceanogr., 52, 3033–3047, https://doi.org/10.1175/JPO-D-21-0263.1.
Xu, G., and Coauthors, 2022: Impacts of model horizontal resolution on mean sea surface temperature biases in the Community Earth System Model. J. Geophys. Res. Oceans, 127, e2022JC019065, https://doi.org/10.1029/2022JC019065.
Yang, P., Z. Jing, B. Sun, L. Wu, B. Qiu, P. Chang, and S. Ramachandran, 2021a: On the upper-ocean vertical eddy heat transport in the Kuroshio Extension. Part I: Variability and dynamics. J. Phys. Oceanogr., 51, 229–246, https://doi.org/10.1175/JPO-D-20-0068.1.
Yang, P., Z. Jing, B. Sun, L. Wu, B. Qiu, P. Chang, and S. Ramachandran, and C. Yuan, 2021b: On the upper-ocean vertical eddy heat transport in the Kuroshio Extension. Part II: Effects of air–sea interactions. J. Phys. Oceanogr., 51, 3297–3312, https://doi.org/10.1175/JPO-D-21-0013.1.
Zhang, J., Z. Zhang, and B. Qiu, 2023: Parameterizing submesoscale vertical buoyancy flux by simultaneously considering baroclinic instability and strain‐induced frontogenesis. Geophys. Res. Lett., 50, e2022GL102292, https://doi.org/10.1029/2022GL102292.