ArticlePDF Available

Submesoscale Ageostrophic Motions Within and Below the Mixed Layer of the Northwestern Pacific Ocean

Wiley
Journal of Geophysical Research: Oceans
Authors:
  • South China Sea Institute of Oceanology, Chinese Academy of Sciences.

Abstract and Figures

Submesoscale dynamics below the mixed layer (ML) and their mechanisms are still unclear. By a series of nested simulations in the Pacific Northwest with high horizontal resolution of ∼500 m, this study reveals that there exist strong submesoscale ageostrophic motions in the upper pycnocline of the Kuroshio Extension region. These motions exhibit enhanced lateral buoyancy gradient and vigorous vertical velocity but with weak vertical vorticity distinct from the ML submesoscale activities. The vertical velocity in the high‐resolution simulation reaches tens of meters per day, consistent with recent observations (e.g., SubMESI and OSMOSIS). Our analysis shows that the enhanced vertical velocity is mostly attributed to the along‐isopycnal motions at the Kuroshio front, but in the region nearby the large vertical velocity mostly arises from the wave‐like vertical movement of isopycnals. To understand the mechanisms for the large vertical velocity, this paper further examined the instability of the flow and the frequency‐wavenumber spectra of vertical vorticity, lateral divergence, lateral buoyancy gradient, and vertical velocity. A criterion based on the ratio between divergence and vorticity variance in spectral space is used to roughly identify the upper bound of unbalanced submesoscales. The results suggest that the high‐frequency, high‐wavenumber processes dominate the vertical motions within and below the ML and significantly enhance the net vertical heat transport between the ML and the ocean interior. This study seeks to provide comprehension of the submesoscale ageostrophic motions below the ML and their impacts on the upper ocean.
This content is subject to copyright. Terms and conditions apply.
1. Introduction
The oceanic flow is known to be highly turbulent and consists of motions on a wide range of scales. With the
great success in satellite altimetry, geostrophic eddies with horizontal scales of a few hundred kilometers can
be routinely observed (Chelton etal., 2007; Fu etal., 2010). The mesoscale features and their impacts on the
transport of oceanic tracers have been well studied in the past decades (e.g., Chelton etal., 2011; Ferrari &
Wunsch,2009; Qiu & Chen,2013; Xu etal., 2016; Zhang etal.,2014). Recently, the physics of submesoscale
turbulence has attracted great attention as it is not well understood due to the difficulty and challenge in observ-
ing submesoscale dynamics (McWilliams, 2016). Recent literature has reported that submesoscale processes
play a crucial role in the forward energy transfer from the geostrophic scale toward dissipation (e.g., Cao
etal.,2021; Capet etal.,2008a; D’Asaro etal.,2011; Jing etal.,2021; Kaneko etal.,2013; Lazaneo etal.,2020;
McWilliams,2016; Thomas etal.,2008; Wang etal.,2018; Zhang etal.,2016) and highlighted their importance
to the vertical transport of heat, salt, and nutrients, which ultimately is important for biological dynamics (e.g.,
Lévy etal.,2012,2001; Mahadevan & Tandon,2006; Rosso etal.,2014; Siegelman etal.,2020; Su etal.,2018).
Abstract Submesoscale dynamics below the mixed layer (ML) and their mechanisms are still unclear. By
a series of nested simulations in the Pacific Northwest with high horizontal resolution of ∼500m, this study
reveals that there exist strong submesoscale ageostrophic motions in the upper pycnocline of the Kuroshio
Extension region. These motions exhibit enhanced lateral buoyancy gradient and vigorous vertical velocity but
with weak vertical vorticity distinct from the ML submesoscale activities. The vertical velocity in the high-
resolution simulation reaches tens of meters per day, consistent with recent observations (e.g., SubMESI and
OSMOSIS). Our analysis shows that the enhanced vertical velocity is mostly attributed to the along-isopycnal
motions at the Kuroshio front, but in the region nearby the large vertical velocity mostly arises from the
wave-like vertical movement of isopycnals. To understand the mechanisms for the large vertical velocity, this
paper further examined the instability of the flow and the frequency-wavenumber spectra of vertical vorticity,
lateral divergence, lateral buoyancy gradient, and vertical velocity. A criterion based on the ratio between
divergence and vorticity variance in spectral space is used to roughly identify the upper bound of unbalanced
submesoscales. The results suggest that the high-frequency, high-wavenumber processes dominate the vertical
motions within and below the ML and significantly enhance the net vertical heat transport between the ML and
the ocean interior. This study seeks to provide comprehension of the submesoscale ageostrophic motions below
the ML and their impacts on the upper ocean.
Plain Language Summary The ocean is usually known as a dynamical system of geostrophic
turbulence. This study uses a high-resolution (∼500m) numerical simulation to investigate the geostrophically
unbalanced motions with horizontal length scales of a few kilometers (submesoscales) in the upper ocean of the
northwestern Pacific. The results show active submesoscale processes (e.g., submesoscale fronts and eddies) at
the surface and even below the mixed layer (ML), which have distinct dynamical features from large-scale or
mesoscale flows. Although these motions are not the primary energy reservoir, they play a significant role in
the energy transfer between scales and can drive much stronger vertical motions than large-scale and mesoscale
processes within and below the ML. This is consistent with the recent observations in the other regional oceans.
Importantly, our results suggest that these motions can drive a significant net heat transport between the ML
and the ocean interior at horizontal length scales much smaller than the baroclinic Rossby deformation radius.
CAO AND JING
© 2022. American Geophysical Union.
All Rights Reserved.
Submesoscale Ageostrophic Motions Within and Below the
Mixed Layer of the Northwestern Pacific Ocean
Haijin Cao1,2 and Zhiyou Jing3
1Key Laboratory of Marine Hazards Forecasting, Ministry of Natural Resources, Hohai University, Nanjing, China, 2College
of Oceanography, Hohai University, Nanjing, China, 3State Key Laboratory of Tropical Oceanography, South China Sea
Institute of Oceanology, Chinese Academy of Sciences, Guangzhou, China
Key Points:
The submesoscale motions below
the mixed layer (ML) show different
features from the mixed-layer
submesoscale processes
The mechanisms for the enhanced
vertical velocity rely on along-
isopycnal motions and internal waves
The high-frequency, high-wavenumber
processes dominate the net vertical
heat transport between the ML and the
ocean interior
Supporting Information:
Supporting Information may be found in
the online version of this article.
Correspondence to:
Z. Jing,
jingzhiyou@scsio.ac.cn
Citation:
Cao, H., & Jing, Z. (2022). Submesoscale
ageostrophic motions within and below
the mixed layer of the northwestern
Pacific Ocean. Journal of Geophysical
Research: Oceans, 127, e2021JC017812.
https://doi.org/10.1029/2021JC017812
Received 21 JUL 2021
Accepted 1 FEB 2022
10.1029/2021JC017812
RESEARCH ARTICLE
1 of 22
Journal of Geophysical Research: Oceans
CAO AND JING
10.1029/2021JC017812
2 of 22
Submesoscale flows are characterized by the O(1) Rossby number,
Ro =
, and Froude number,
Fr =𝑁
.

is the relative vorticity, f is Coriolis frequency, V is a characteristic velocity with a horizontal length scale of
L, N is the buoyancy frequency, and h is the vertical length scale (McWilliams,2016). Physically, submesoscale
processes differ from geostrophic flows by the intensification of ageostrophic motions but are still influenced by
Earth rotation, manifested by the submesoscale horizontal buoyancy gradient, divergence/convergence, straining,
and large vertical vorticity (Capet etal.,2016; Shcherbina etal.,2013).
In the open ocean, it is difficult to capture intermittent and fast-evolving submesoscales through traditional
cruise surveys (McWilliams,2016; Shcherbina et al., 2013). Although surface submesoscale eddies or fronts
are frequently detected from high-resolution remote sensing images (e.g., Liu etal.,2015; Yu etal.,2018; Zeng
etal.,2014; Zheng,2017), the dynamical link between the surface and the oceanic interior is much more complex
for submesoscale flows than larger-scale geostrophic flows. It is a great challenge to reconstruct the submesoscale
features in the ocean interior based on the sea surface height (Qiu etal.,2016,2020). Instead, high-resolution
numerical simulations, together with satellite data, have been widely used to reproduce the submesoscale phys-
ics (e.g., Balwada etal., 2018; Capet etal., 2008b; Gula etal.,2014; Klein etal.,2008; Lapeyre etal.,2006;
Qiu et al., 2014; Rocha etal.,2016) and to explore their generating mechanisms (Fox-Kemper etal.,2008;
McWilliams,2017; McWilliams etal.,2015; Srinivasan etal.,2019).
Previous studies focus mainly on submesoscale processes within the mixed layer (ML). However, recent stud-
ies suggest that the ocean interior is not always in quasi-geostrophic balance with small Rossby number, but
with enhanced vertical motions at submesoscales (Siegelman etal.,2020; Yu etal.,2019a; Zhang etal.,2021).
Case studies investigated the ageostrophic dynamics during restratification (Johnson etal.,2020a,2020b). These
motions are particularly energetic in strong frontal systems (e.g., the Kuroshio front) and can drive considera-
ble vertical flux in the ocean interior. However, to date, their dynamical mechanisms remain unclear. Based on
a series of nested simulations in the eddy-rich northwestern Pacific, this study investigates the submesoscale
features, with a focus on the geostrophically unbalanced scales (high-frequency, high-wavenumber), highlighting
their significant contribution to vertical velocity within and below the ML. Here, submesoscale ageostrophic
motions refer to the motions that can drive large vertical velocity at high frequencies and high wavenumbers (e.g.,
larger than the Coriolis frequency f and the wavenumber of 1×10
−4cpm). The outline of this paper is as follows.
The next section briefly describes the simulations. Section3 compares the submesoscale characteristics (relative
vorticity, divergence, buoyancy gradient, strain field, and vertical velocity) of the simulations at low, middle,
and high resolution (referred to as LR, MR, and HR). In this section, the dynamical regimes of the large vertical
velocity are investigated in detail. Section4 further investigates the scale range of ageostrophic motions and the
dynamical mechanisms for the large vertical velocity. Finally, the results are summarized in Section5.
2. Model Description
A nested simulation suite in the northwestern Pacific was conducted using the Regional Oceanic Modeling
System (ROMS; Shchepetkin & McWilliams,2005). Three-layer offline nesting was applied from a coarse hori-
zontal resolution of ∼7.5km to the intermediate resolution of ∼1.5km and finally to the highest resolution of
∼0.5km (Figure1). The ∼7.5-km simulation covers a large domain from 10°S to 45°N and from 95°E to 170°E
with a 20-year spin-up to reach a statistically steady state before starting the one-way nesting simulations. The
next simulation in the nesting hierarchy (∼1.5km) extends from 28°N to 43.5°N and from 138°E to 162°E and
ran for a whole year to provide boundary and initial conditions for the higher resolution simulation. The final
∼0.5-km simulation is run in a smaller domain: 30°–40°N, 142°–155°E for 6weeks (after 4-week spin-up, the
2-hourly outputs of last 2weeks from April 28 to May 12). Since the resolved submesoscale structures are statis-
tically quasi-steady, we select snapshots at some time instants to focus the analysis. Each of the simulations (LR,
MR, and HR) is run on a curvilinear, latitude-longitude grid, and terrain-following S-coordinates of 60 vertical
levels. The vertical layers are refined with 24 layers over the upper 200m.
The boundary conditions and initial state for the LR simulation are provided by the monthly averaged Simple
Ocean Data Assimilation (SODA) ocean climatology data (Carton & Giese,2008). The surface atmospheric forc-
ing such as wind stress, heat, and freshwater fluxes is provided by the Quick Scatterometer (QuikSCAT) data set
and the International Comprehensive Ocean Atmosphere Data Set (ICOADS; Woodruff etal.,2011). The K-pro-
file parameterization (KPP) is used for the subgrid vertical mixing of momentum and tracers (Large etal.,1994).
Journal of Geophysical Research: Oceans
CAO AND JING
10.1029/2021JC017812
3 of 22
Note that we used the climatological daily wind forcing and did not include tidal forcing in the simulations. The
simulations on regional circulation, ML depth, sea surface temperature, and kinetic energy have been validated
against satellite measurements and available in situ observations (see the Supporting Information file and model
description of Luo etal.(2020), Huang etal.(2020), and Cao etal.(2021). The comparison with the measure-
ments shows that the simulations have reached a steady state after 20-year spin-up and are sufficiently accurate
to characterize the climatological Pacific conditions and delineate local submesoscale features. In the HR simu-
lation, the computational domain shows a shallow ML depth of about 35m using the criteria of a density differ-
ence of 0.03kgm
−3 from the surface layer, consistent with some observations (de Boyer Montégut etal.,2004;
Kara etal., 2003). As such, the simulated period in this study exhibits less active submesoscales compared to
the wintertime (Sasaki etal.,2014). The output shows a robust set of statistics of submesoscale characteristics,
so the randomly selected snapshot can nearly represent the submesoscale features within a narrow time window.
3. Submesoscale Features
3.1. Comparison of Submesoscale Characteristics (LR, MR, and HR)
The submesoscale features in the LR, MR, and HR simulations are compared at depths of z=−10 m, −50m,
and −200m, representing the upper ML, the base of the ML, and the pycnocline, respectively. A comparison
of kinetic energy from different simulations suggests differential meandering shapes (Figure S1 in Supporting
InformationS1), as some of the subgrid processes are parameterized in the LR and MR simulation but resolved
in the HR simulation. The submesoscale eddies can modify the large-scale and mesoscale flows by inverse kinetic
energy cascade (Schubert etal.,2020).
Figure 1. (a) The model surface temperature in domains of the nested models at horizontal resolutions of ∼7.5km (low resolution (LR)), ∼1.5km (middle resolution
(MR)), and ∼0.5km (high resolution (HR)), respectively. The dashed-line box denotes a subregion for the analysis of vertical velocity in Section3.4. (b) Snapshot of
simulated upper-ocean temperature field at HR.
Journal of Geophysical Research: Oceans
CAO AND JING
10.1029/2021JC017812
4 of 22
The O(1) Ro (

) is generally used to characterize submesoscale flows (Bachman etal.,2017; Gula etal.,2014;
Thomas etal.,2008). As shown in Figure2 (a snapshot at a randomly selected time, 10:00 on 30 April), all the
vorticity fields show a horizontally elongated line pattern with large positive/negative values of relative vorticity
along the Kuroshio jet, which is a sign of sharp frontogenetic buoyancy fronts (Hoskins & Bretherton,1972;
McWilliams, 2016). The HR simulation with the highest horizontal resolution compared to the LR and MR
simulations shows a denser population of eddies at the three depths (the right column compared to the left and
middle columns in Figure2). A comparison of the snapshots for different depths (i.e., Figure2, right column)
shows a clearly weaker vorticity field at the 200-m depth than at the 10-m and 50-m depths. The 200-m depth is
mostly below the ML throughout the year. Those small eddies certainly have limited vertical scales and cannot
extend to depth (D'addezio etal.,2020; Liu etal.,2020)—submesoscale eddies are mostly contained in the ML
(Figure 2, right column). This also restrains the development of submesoscale eddies to be local (Boccaletti
etal.,2007; Fox-Kemper etal.,2008). Similar to the vertical vorticity, horizontal divergence,

=
+
, and
Figure 2. Snapshots of relative vorticity normalized by

at different depths (z=−10m, −50m, and −200m) from different simulations (LR, MR, and HR) at a
randomly selected time, 10:00 on 30 April. LR is short for low resolution (left column), MR is short for middle resolution (middle column), and HR is short for high
resolution (right column).
Journal of Geophysical Research: Oceans
CAO AND JING
10.1029/2021JC017812
5 of 22
strain rate,

2
, are also largely intensified with smaller scales in the HR simulation
than in the MR and LR simulation, which is indicative of the development of ageostrophic motions (Figures S2
and S3 in Supporting Information S1). As shown in Figure S4 in Supporting InformationS1, the divergence
is closely related to the strain-induced frontogenesis (Barkan etal.,2019). Although the strain is reduced with
depth, it remains strong near the Kuroshio jet and can drive ageostrophic motions, suggesting the generation of
ageostrophic motions below the ML.
Figure3 shows the magnitude of lateral buoyancy gradient (
||
), where
=(1−0)
is the buoyancy, with
g the gravitational acceleration,

the potential density, and

0
a reference density of 1,025kgm
−3. Compared
with the LR and MR simulation, the HR simulation displays more active fronts at smaller scales. Stronger
submesoscale buoyancy gradients appear more frequently on the north side of the Kuroshio jet (right column
of Figure3). In the ML with weak stratification, the strain-induced frontogenesis can drive sharp density fronts
Figure 3. Snapshots of lateral buoyancy gradient at different depths (z=−10m, −50m, and −200m) from different simulations (LR, MR, and HR) at a randomly
selected time, 10:00 on 30 April. LR is short for low resolution (left column), MR is short for middle resolution (middle column), and HR is short for high resolution
(right column).
Journal of Geophysical Research: Oceans
CAO AND JING
10.1029/2021JC017812
6 of 22
(McWilliams,2016; Thomas etal.,2008). Meanwhile, a part of the fronts/filaments can transform into submesos-
cale eddies through ML baroclinic instabilities (Boccaletti etal.,2007; Fox-Kemper etal.,2008). All of these
submesoscale features can result in large

,

, and

at the 10-m depth. Notice that the submesoscale fronts
captured by sheared flows at z=−50m remain energetic. There still exist strong submesoscale density fronts
well below the ML (e.g., z=−200m). The consequence for the submesoscale fronts is discussed in the following
sections.
3.2. Statistics of Vorticity, Divergence, Strain Rate, and Buoyancy Gradient
This section evaluates the statistics of vertical vorticity, lateral divergence, strain rate, and lateral buoyancy gradi-
ent for different simulations at z= −10 m and for different depth levels from the HR simulation (Figure4).
The probability density functions (PDFs) calculated from the HR domain show that higher horizontal resolu-
tion presents greater submesoscale characteristics. Table1 lists the mean, skewness, and standard deviation of
the vorticity, divergence, and strain rate for the simulations. The HR compared to the MR and LR simulations
presents larger skewness for all quantities (Figure4, upper row), as more submesoscale cyclones are resolved
with higher horizontal resolution. Strong anticyclones are rarely seen because the flow is centrifugally unstable
when

is smaller than −1. The distribution of relative vorticity,

, from the HR simulation is markedly
Figure 4. Probability of density function (PDF) of (a, e) normalized vertical vorticity, (b, f) lateral divergence, (c, g) lateral strain rate, and (d, h) buoyancy gradient for
three simulations (upper row, z=−10m) and for three depth levels (bottom row, the high resolution (HR) simulation), respectively.



Mean Sk ew. St. d. Mean Sk ew. St. d. Mean Skew St. d.
LR 0.000 0.650 0.250 −0.001 −0.167 0.087 0.204 1.691 0.160
MR −0.001 2.952 0.437 0.002 −1.951 0.122 0.312 5.527 0.327
HR −0.005 3.988 0.606 −0.002 −1.522 0.187 0.436 7.021 0.460
Table 1
The Mean, Skewness (Skew.), and Standard Deviation (St. d.) of

,

, and

for the Simulations (z=−10m)
Journal of Geophysical Research: Oceans
CAO AND JING
10.1029/2021JC017812
7 of 22
asymmetric with a large positive skewness of 3.988 and a standard deviation of 0.606, which decreases to 2.952
and 0.437 in the MR simulation, and to 0.650 and 0.250 at the 200-m depth, respectively (Table1).
Table2 compares the mean value, skewness, and standard deviation of the variables for the three depth levels,
where the statistical estimations are based on the HR simulation (Figure4, bottom row). The

ranges from
−1 to 2, −0.8 to 1.2, and −0.8 to 0.9 at z=−10m, −50m, and −200m, respectively. Vertically, the

is more
skewed at the surface compared to the greater depths (skewness of 2.679, 1.904, and 0.932 at the depths of 10, 50,
and 200m, respectively) and shows decreasing standard deviation (0.537, 0.417, and 0.275 for the three depths).
They agree with the result shown in Figure2 that there are more submesoscale eddies in the ML than below
and large vorticity values tend to be cyclonic (the positive tail with

> 1), consistent with the observations
(Rudnick,2001; Shcherbina etal.,2013). Below the ML, the vorticity distribution becomes more Gaussian-like
in the absence of topography. The distributions of lateral divergence are fairly symmetric at all depths and become
less skewed with increasing depth. The PDF of the strain rate is more narrowly distributed, with continuously
decreasing peak modes from
=0.175
at z=−10m to
=0.1
at z=−200m. Statistically, high values
of
||
greater than 1×10
−7s
−2 appear more frequently at z= −50m compared to z= −10m and −200m
(Figure4h) because of the variation of the ML base. Distinct from the other parameters analyzed here,
||
is
not monotonically decreasing with depth. Similar results were also reported in observation-based studies (e.g.,
Thompson etal.,2016).
3.3. Wavenumber Spectrum of Vorticity, Divergence, Strain Rate, and Buoyancy Gradient
The wavenumber spectrum analysis is an effective approach to identify the features of submesoscales as functions
of scales. Dynamically, the divergent KE is primarily attributed to unbalanced motions including both ageo-
strophic motions and internal gravity waves (hereafter referred to as IGW), while the spectrum of vertical vortic-
ity indicates the enstrophy, known as the geostrophically balanced part of KE. Here, “ageostrophic” motions refer
to the unbalanced part of submesoscale processes. The partition between balanced and unbalanced motions is
illustrated in the theoretical decomposition (Helmholtz decomposition), whereby the KE is decomposed into the
rotational and divergent components (Bühler etal.,2014; Torres etal.,2018) as follows:
=r+d=−+,=r+d=+,
(1)
where

r
is the rotational component and

d
is the divergent component of

,

r
is the rotational component and

d
is the divergent component of

,

is the stream function, and

is the potential. Then, the Laplacian of stream
function and potential is
Δ=,Δ=+
(2)
Thereby, in the spectral space, the power spectrum of normalized squared vertical vorticity (
RV =22
) can
be regarded as the vortex contribution to KE (
RV
()=
2
2KEv(
)
, in which k is the wavenumber and
KEv()
is
the vortex component of KE spectrum). The transformation of the isotropic 2D to 1D spectrum can be found in
the Appendix of Cao etal.(2019). Thus, a k
−1 vorticity spectrum corresponds to a k
−3 KE spectrum (geostrophic
prediction). Similarly, the spectrum of lateral divergence (
DIV =22
) signifies the divergent component of
KE (
DIV
()=
2
2KED(
)
).
Figure5 compares the spectra from different simulations but within the same spatial domain of the HR simulation
(z=−10m for the left column). The HR simulation shows a flattened vorticity spectrum (a slope of ∼−0.5) at



Mean Sk ew. St. d. Mean Sk ew. St. d. Mean Sk ew. St. d.
10m 0.001 2.679 0.537 0.001 −1.041 0.137 0.290 5.172 0.282
50m 0.002 1.904 0.417 0.000 −0.081 0.111 0.215 3.532 0.162
200m 0.002 0.932 0.275 0.000 −0.020 0.091 0.134 3.022 0.103
Table 2
The Mean, Skewness (Skew.), and Standard Deviation (St. d.) of

,

, and

for Different Depth Levels (HR
Simulation)
Journal of Geophysical Research: Oceans
CAO AND JING
10.1029/2021JC017812
8 of 22
high wavenumbers, indicative of submesoscale flows. The slope of the vorticity spectrum indicates the strength
of submesoscale processes (Figure5; right column). The vorticity spectrum for the 10-m depth is flattened to
k
−0.5 at the submesoscale, whereas the spectral slopes are approximately at −1 for the 50-m and 200-m depths,
indicative of geostrophic turbulence (Figure5e). This suggests that submesoscale eddies are mostly generated
in the ML and have limited vertical scales. Interestingly, the LR divergence spectrum shows the highest peak
near 50km (geostrophic divergence) compared to the MR and HR spectra (Figure5b). It suggests that the diver-
gence is redistributed over scales when submesoscales are resolved. Essentially, the generation of submesoscale
horizontal divergence is dynamically complicated, e.g., in turbulent thermal wind balance (Barkan etal.,2019).
Besides, the spectra of horizontal divergence peak at the submesoscale for all the depths (Figure5f), demonstrat-
ing an enhancement of unbalanced motions arising from submesoscale divergence/convergence. The subpeaks
of divergence spectra at the mesoscale indicate the scale of geostrophic divergence. The lateral strain rate is also
reduced at the 50-m and 200-m depths (Figure5g). The spectra of buoyancy gradient at depths (z=−50m and
Figure 5. Wavenumber spectra of (a, e) normalized vertical vorticity, (b, f) lateral divergence, (c, g) lateral strain rate, and (d,
h) horizontal buoyancy gradient for three simulations (left column, at z=−10m) and for three depth levels (right column, the
high resolution (HR) simulation), respectively. PSD is short for power spectrum density.
Journal of Geophysical Research: Oceans
CAO AND JING
10.1029/2021JC017812
9 of 22
z= −200m) are as energetic as that at 10-m depth until 10
−5cpm (100 km) and become less energetic at the
submesoscale (Figure5h). Note that all the spectra experience a dramatic roll-off near the Nyquist wavenumber
(0.075×10
−3, 0.322×10
−3, and 1×10
−3cpm for each simulation respectively), which is a numerical artifact.
3.4. Vertical Velocity and Vertical Heat Flux
Vertical velocity (w) associated with submesoscale flow fields is typically larger than that of the mesoscale fields
(Calil & Richards,2010; Lévy etal.,2012; Mahadevan & Tandon,2006; McWilliams,2016). A comparison of
the vertical velocity between the simulations shows enhanced instantaneous w in the HR simulation, suggesting
that the horizontal resolution is a sensitive parameter for the reproduction of vertical motions driven by ageo-
strophic processes with smaller length scales (Figure6). Here, we zoom into a subregion marked by the dash-line
box in Figure1 for removal of bottom effects like seamounts. Differing from the vertical vorticity or horizontal
divergence, w shows larger amplitude at the 200-m depth instead of at the shallower depths. The w at 10-m
depth concentrates on the filamentary structures that are elongated by the strain along the jet; while the w at the
200-m depth shows larger amplitude and spreads over a wider extent. Some regions of the w exhibit small-scale
crisscross structures which are indicative of internal waves. This implies that the large w could partly result from
the spontaneous emission of inertial gravity waves near the Kuroshio front (Danioux etal.,2012; Plougonven &
Figure 6. Snapshots of vertical velocity at different depth levels (z=−10m, −50m, and −200m) from different simulations (low resolution (LR), middle resolution
(MR), and high resolution (HR)) at 10:00 on 30 April.
Journal of Geophysical Research: Oceans
CAO AND JING
10.1029/2021JC017812
10 of 22
Snyder,2007). While the strong stratification below the ML tends to suppress the growth of submesoscale insta-
bilities but indeed supports the propagation of internal waves.
Unlike the large vertical velocity in the ML which has been shown to be related to submesoscale dynamics such
as ML instabilities and frontogenesis (Fox-Kemper etal.,2008; McWilliams etal.,2017), the mechanism lead-
ing to the large w in the pycnocline of the Kuroshio front is not yet identified. The large w is likely associated
with either internal waves or motions along isopycnals (Klymak etal., 2016). Symmetric instability may also
partly contribute to the vertical motions (Thomas etal.,2013). To better understand the vertical velocity in the
pycnocline, Figure7 shows the 3D snapshots of vertical vorticity, vertical velocity, and temperature field on the
density surface of σθ=25.8. In this way, we can clearly identify the location of the density front and its sharpness.
More importantly, the temperature variances on the isopycnal are also visualized (Figure7b), as an indication
of along-isopycnal motions. The vorticity filed on the σθ=25.8 isopycnal, which extends from near-surface to
Figure 7. (a) Vertical vorticity, (b) temperature, and (c) vertical velocity on the density surface σθ=25.8 for the subregion. (d) Vertical velocity w, (e) the estimated
wiso, and (f) the estimated ratio wiso/w based on the scaling method of Freilich and Mahadevan(2019) at 200-m depth. In (e) and (f), the locations where the scaling
relationship is negative because of the negative buoyancy frequency (N
2) are blanked.
Journal of Geophysical Research: Oceans
CAO AND JING
10.1029/2021JC017812
11 of 22
500-m depth, exhibits a few eddy-like and filamentary structures (Figure 7a). Figure7c shows strong vertical
velocity with small length scales, especially at the sloping front. Note that the interpolation of vertical velocity
to the density surface may not be accurate because of the limited vertical layers of the model output especially at
depth. Additionally, the method presented by Freilich and Mahadevan(2019) is employed to decompose the verti-
cal velocity w into two components: wiso along the sloping isopycnal surfaces and wlift as the lifting of isopycnal
surfaces. The scaling relationship is as follows:
iso
(
2
2
),
(3)
where

2
2
is the isopycnal slope with

2
=
|
|
and

2
=
and

is the inverse aspect ratio esti-
mated as

=
()2+()2
()2+()
2
. To illustrate, we apply the decomposition to
the vertical field at 200-m depth to examine the contribution of the two components (Figure7, right column).
The wiso dominates the vertical velocity along the Kuroshio front, suggesting that the motions at sloping isop-
ycnal surfaces provide an important pathway for the vertical transport of tracers between the surface and the
ocean interior (Mahadevan etal.,2020). We note that the large vertical velocity, which is observed in the other
parts of the computational domain, is likely high-mode internal waves emitted by frontal processes (Shakespeare
& Taylor, 2015). In Figures7e and 7f, the locations where the scaling relationship is negative are blanked out
because of the negative buoyancy frequency (N
2).
To gain more insight into the dynamical regimes for the large vertical velocity, the spatial correlations of vertical
velocity with normalized vertical vorticity, buoyancy gradient, and frontogenesis function are examined. The
frontogenesis function is defined as
s
=−
𝑏
(4)
where
is the frontogenetic vector that can be expressed as (Hoskins etal.,1978)
𝐐
=
𝑢𝑥𝑣𝑥
𝑢𝑦𝑣𝑦
𝑏𝑥
𝑏𝑦
.
(5)
Fs indicates the evolution of a buoyancy gradient, i.e., frontogenesis (positive Fs) or frontolysis (negative Fs).
The subscripts, x and y, denote the gradient in the zonal and meridional direction, respectively. In Figure8, we
perform a linear fit between vertical velocity and the normalized vorticity as well as the horizontal buoyancy
gradient and frontogenesis function. We also include the correlation coefficient and 95% confidence interval
to demonstrate the fidelity of the linear relationship. A high correlation coefficient typically indicates the loca-
tions of large |w| correspond well with large |

|,
||
, or Fs. In Figure8a, it is interesting that the |w| shows
a clear trend with
||
at 10-m depth (the blue dots), indicating that the large values of vertical velocity are
Figure 8. Plots showing the trend of |w| against |

|, buoyancy gradient, and frontogenesis function from the subregion of high resolution (HR) simulation. Blue,
red, and green dots indicate different depth levels representing the mixed layer (ML), the base of the ML, and the upper thermocline, respectively. Error bars of 95%
confidence interval are also plotted.
Journal of Geophysical Research: Oceans
CAO AND JING
10.1029/2021JC017812
12 of 22
related to the submesoscale eddies in the ML. The correlation coefficient
at the 50-m depth is 0.54 (weak correlation), much lower than that of the
near-surface layer. However, at the 200-m depth, a negative slope between
|w| and
||
is shown within a limited range of
||
(Figure8a, the green
dots). In contrast, |w| shows a positive correlation with the buoyancy gradient
(
||
) at all three depths (Figure8b). The large vertical velocity events below
the ML correlate better with the large buoyancy gradient than with the large
vertical vorticity, although the error bars for high values of
||
at z=−50m
and −200m are relatively large. The enhanced vertical velocity is a result of
frontal effects. Furthermore, the high correlation between |w| and Fs indi-
cates that the high values of vertical velocity are induced by frontogenesis/
frontolysis at the three depths (Figure8c). The frontogenetic effects tend to
decrease with smaller Fs at depths (Figure8c, the red and green dots) but are
associated with larger vertical velocity |w|. In this context, the frontogenetic
processes are of great importance to the large vertical velocity below the ML,
driving efficient vertical transport especially at the site of the sharp front
(recall Figure7).
Although the results show evidence of large vertical velocity, it is unknown
whether these motions can drive strong net vertical transport. In Figure 9,
the statistics of the HR vertical velocity separates the upward (positive w+)
and downward (negative w) velocities. To evaluate the contribution of large
vertical velocities (
||
30
m/day), the ratios between the large positive and
negative vertical velocity in Figure9 (w/w+ within the range of
||
30
m/
day) are calculated to be 1.91 for the 10-m depth, 1.13 for the 50-m depth, and
1.12 for the 200-m depth, respectively. This is consistent with previous stud-
ies that indicate stronger downwelling along ageostrophic secondary circula-
tions (e.g., McWilliams,2017). In addition, this also suggests that the vertical
transport tends to be downward with more asymmetric distribution in the ML
than below, in a way, illustrating the efficiency in vertical transport at differ-
ent depths. Vertical heat flux (VHF) is also estimated as
VHF =0p
,
where

p
= 3,985 Jkg −1K
−1 is the heat capacity of seawater, and

and

refer to the anomalies of vertical velocity and temperature by removing the
domain-averaged values. The positive or negative VHF indicates upward or
downward heat flux, respectively. The time and domain-averaged VHF in the
upper 200m are plotted in Figure10 (the solid lines). The VHF experiences
a quick increase from the surface to the middle of the ML (∼25m), reach-
ing peak values of 86.3W/m
2 (1-day mean), and 83.8W/m
2 (3-day mean),
respectively. Below the ML, the VHF remains positive, illustrating the net
upward vertical heat transport. As linear IGW cannot drive clear VHF, there
should be some other processes contributing to the enhanced net vertical heat
transport. Here, a low-pass filter (>10km) was applied to filter out the effect
of high-wavenumber motions for

and

. In contrast, the low-pass vertical
profiles show a decrease in the ML and almost reduce by half below the ML
(the dashed lines in Figure 10). It follows that the VHF in the ML mostly
relies on the high-wavenumber submesoscale processes, while below the ML,
it is only partly provided by these processes. Note that the possible impacts of
nonlinear IGW and their interaction with background flows are not excluded.
4. Analysis of Dynamics and Regimes
The above results have shown the submesoscale features characterized by
strong ML eddies with weaker vertical velocity in the ML and by active
fronts/filaments with stronger vertical velocity below the ML. However, the
Figure 9. Probability of density function (PDF) of vertical velocity at three
depths for the high resolution (HR) simulation (blue line for 10-m depth,
red line for 50-m depth, and yellow line for 200-m depth).
||
=30m/day is
marked by the dashed lines.
Figure 10. Time and domain-averaged vertical heat flux (VHF) over the
subregion (Figure6) and 1day (30 April; the blue line) and 3days (30 April
to 2 May, the red line). The dashed lines are the corresponding low-pass VHF
derived with low-pass

and

(>10km).
Journal of Geophysical Research: Oceans
CAO AND JING
10.1029/2021JC017812
13 of 22
dominant process accounting for the enhanced vertical velocity in the ocean
interior remains unclear. Hence, this section further examines the flow state
with the Okubo-Weiss (OW) parameter and investigates the instability condi-
tions by analyzing the Ertel potential vorticity and the frequency-wavenum-
ber spectrum.
4.1. Okubo-Weiss Parameter
The upper-ocean submesoscale turbulence is generated by strain field
and horizontal buoyancy gradient in the form of interactive submesoscale
fronts/filaments and eddies. For example, filaments can break into a train
of submesoscale vortices by shear instabilities (Gula etal., 2014). In this
section, the OW parameter,
OW
=
2
2
, is employed to diagnose the state
of the fast-evolving submesoscale flows. Here, the sign of OW determines
the behavior of a tracer gradient (e.g., buoyancy gradient)—positive OW
indicates a growth of tracer gradient induced by strain rate, while negative
OW suggests that the flow tends to be eddying (Gula etal.,2014; Hua &
Klein, 1998). The joint probability distribution functions (JPDF) of verti-
cal vorticity and lateral strain rate are presented in Figure11. The statistical
result shows that the percentages of positive OW for z=−10m, −50 m,
and −200m are 64.8%, 68.3%, and 69.1%, respectively. This result indicates
that the depths show stronger growth of buoyancy gradient in a straining
flow regime with stronger Fs (Figure8c), which can drive stronger vertical
flux (Nagai etal.,2015). Also, the JPDF at z=−10m shows that the large
Ro (Ro> 1) occurs near the
||
=
(shear flow) line and correlates well
with the strain rate. In the ML, the submesoscale eddies and frontogenesis/
filamentogenesis are of comparable importance, as the weak stratification is
favorable for the ML baroclinic instabilities and strain-induced frontogenesis
(Callies etal., 2016). At z= −200m, both the vorticity and strain rate are
reduced, so the JPDF spreads and more positive OW occurs, indicative of the
growth of tracer gradient by flow strain.
4.2. Ertel Potential Vorticity
The Ertel potential vorticity (PV) is estimated for examining the instabilities
associated with submesoscale processes in the rotational and stratified flow.
In the diagnostic estimations, we assume that the large-scale mean flows are
to the leading order in geostrophic balance. The Ertel PV can be expressed as
=(+)+(×)
h𝑏
(6)
where the horizontal gradients of vertical velocity are negligible. Then the Ertel PV can be decomposed into
two components: the vertical component,

vert
=(+)
, and the baroclinic component,

 =(×)
h
.
The sign of Ertel PV indicates the flow state (negative PV in the northern hemisphere corresponds to instabili-
ties; Thomas etal.,2013). Here, the two components of Ertel PV are analyzed separately. As in Equation6, the
negative

can arise from unstable stratification, anticyclonic lateral shear, or a negative baroclinic component
(Thomas etal.,2013). In comparison, the vertical component tends to be the dominant constituent for the Ertel
PV in most of the computational domains except near the Kuroshio front (Figure11).
The mesoscale and submesoscale features are highly heterogeneous throughout the computational domain, with
denser signatures near the Kuroshio front (Figure11, snapshots of Ertel PV on 30 April for three depth levels).
At z=−10m, the qve rt is small in the southern flank, indicative of reduced PV for ongoing instabilities; while
the strong vertical stratification near 40°N (large positive qvert) increases the Ertel PV, acting to stabilize the flow.
Differently, the baroclinic component, qbc, exhibits stronger and denser filamentous structures in the north than
in the south. In particular, the strong negative qbc along the Kuroshio front makes q<0 at the near-surface layer,
Figure 11. Joint probability distribution functions of vorticity versus strain
rate at three depths. The probability is in logspace (log10).
Journal of Geophysical Research: Oceans
CAO AND JING
10.1029/2021JC017812
14 of 22
which is favorable for instabilities. This pattern coincides with the distribution of lateral buoyancy gradient (recall
Figure3), arising from submesoscale frontogenetic processes. At z=−50m, the qvert is mostly positive indicat-
ing stable stratification (except for the stream region between 33.5°N and 36°N). The ML base is highly varying
especially near eddies, similar to the observational results (Thompson etal.,2016). This could partly explain the
spatial heterogeneity of the qbc field at the ML base (z=−50m). The depth of 50m can be sometimes within
the ML with small qvert or within the pycnocline with large qvert. Additionally, the ML of this region is quickly
shoaling during the springtime (de Boyer Montégut etal.,2004; Kara etal.,2003) when there should be enhanced
ML restratification (Fox-Kemper etal.,2008). The local restratification may cause submesoscale ageostrophic
motions to restore geostrophic balance. All these processes are particularly active near the Kuroshio front, i.e.,
these ageostrophic processes likely stem from the frontal instabilities at mesoscale and submesoscale (Johnson
etal.,2020a) and can extend to the deep along the frontal isopycnals. At the 200-m depth, although submesoscale
instabilities are largely suppressed (mostly positive Ertel PV for stable flow), there still exist negative qbc with
filamentous structures. Compared to the 50-m depth, the 200-m depth shows more pronounced signatures of
mesoscale eddy rings at 142.3°E, 34.2°N and 145.5°E, 38.0°N (Figure12, middle column). Below the ML, the
baroclinicity in the periphery of the mesoscale eddies is enhanced, which is likely the causation of the enhance-
ment of vertical motions.
The frontal instability is further examined using an improved instability criterion (Buckingham etal., 2021a,
2021b), which introduces a nondimensional number Cu into the classical criterion that uses the Ertel PV q multi-
Figure 12. Snapshots (high resolution (HR)) of Ertel PV being decomposed into vertical (qvert, left column) and baroclinic (qbc, middle column) components for
different depth levels (z=−10m, −50m, and −200m) at 10:00 on 30 April. The solid line in the first plot indicates the cross-front section analyzed in Figure13.
Journal of Geophysical Research: Oceans
CAO AND JING
10.1029/2021JC017812
15 of 22
plied by the Coriolis parameter f ((1+Cu)fq<0 as a necessary criterion for instability. Here, we take a cross-
front section between 151.09°E, 37.24°N and 151.47°E, 36.75°N, e.g., to investigate the instability conditions in
the upper ocean (marked by the solid line in the first plot of Figure12). For curved fronts, the discriminant for
possible instability can be described as
= (1 + Cu)(1 + Ro) (1 + Cu)2
Ri
−1 <0,
(7)
where Cu, Ro, and Ri are the curvature, Rossby, and Richardson number. 1+Cu is defined as the nondimensional
absolute angular momentum. Cu can be calculated as
Cu
=
2
𝑉
𝑓𝑟 ,
(8)
where
𝑉
is the along-front flow velocity. r is the radius of curvature estimated as
= 1∕
where
=
𝑥 𝑥
(
2
+
2
)
3∕2
(9)
is the geometric curvature, and
𝐴
,
𝐴
,
𝐴
, and
𝐴
denote the first and second derivatives of zonal and meridional
displacements along the frontal boundary (see Appendix B of Buckingham etal.,2021a for detail). Here, the
smoothed contours of temperature are used to calculate the radius of curvature. In Equation7, the discriminant
would be negative when the parameter (1+Cu) is large for small Richardson number. The
|
|
, Ri, 1+ Cu,
and

along the cross-front section are shown in Figure13. The Richardson number at the sharp front remains
small (about 3) even at the 200-m depth (Figure13b). As mentioned above that the Rossby number may not be
an important indicator for submesoscale activities in the ocean interior, the Richardson number still plays an
important role in reducing the stability discriminant

(Equation7). In Figure13c, the nondimensional absolute
angular momentum (1+ Cu) tends to be negative along the front, resulting in negative discriminant for insta-
bilities (

< 0). In this discriminant, the instability would occur when 1+ Cu is negative even if Ro is small.
Figure 13. (a) Lateral buoyancy gradient (
||
), (b) the estimated Richardson number in logspace (log10Ri), (c) the nondimensional absolute angular momentum
(1+Cu), and (d) the discriminant for possible instability (

<0
) along the cross-front section.
Journal of Geophysical Research: Oceans
CAO AND JING
10.1029/2021JC017812
16 of 22
According to the definition for instability categories (see Appendix B of Buckingham etal.,2021a), there would
occur symmetric instability at the cross-front section even below the ML, accounting for the large along-isopyc-
nal vertical velocity (Thomas etal.,2016; Yu etal.,2019b).
4.3. Frequency-Wavenumber Spectrum Analysis
Submesoscale and IGW signatures are shown in the simulation. However, these processes (e.g., frontogene-
sis, ML instabilities, and IGW) are usually concurrent with overlapped spatial and time scales and cannot be
disentangled easily in the realistic simulation. In particular, the IGW within the stratified layers can result
in strong horizontal buoyancy gradient and vertical velocity, although the spontaneous IGW are underesti-
mated in this simulation without the forcing of tide and high-frequency wind (Vanneste, 2013). Here, the
frequency-wavenumber spectrum is employed to identify the contribution to vertical vorticity, horizontal
buoyancy gradient, and vertical velocity in different spatial and time scales (Figure14; the dispersion rela-
tion curves of linear IGW are marked in solid curves). The variance spectra of vertical vorticity (

) and
horizontal buoyancy gradient (
||
) for three depths show a similar shape and are predominately contrib-
uted by the high-wavenumber, high-frequency motions. No clear signature of low-mode IGW is shown in
the variance spectra of

and
||
(left and middle column, Figure14). In contrast, the variance of vertical
velocity (w) shows some signatures of IGW at the 200-m depth (Figure14i). Interestingly, the submesoscale
w variance is increased with the increase of depth (Figures14c, 14f, and14i), suggesting enhanced vertical
motions below the ML. However, these motions only lead to a limited enhancement of net vertical heat
transport (recall Figure10). One possible explanation is that the vertical motions at depths partly result
from high-mode IGW, which do not significantly contribute to net vertical transport. The high-mode IGW
near the Kuroshio front should play an important role in vertical motions, though no high-frequency wind
forcing is used in this simulation.
Given the basic definition for balanced and unbalanced KE in the spectra (Section3.3), the frequency-wave-
number spectra can also be used to partition the balanced and unbalanced motions and to identify their scale
range. Here, a ratio between the divergence and vorticity spectrum (
=divergencevorticity
) is defined. If
R1
, the flow is considered to be rotational (quasi-geostrophic). The frequency-wavenumber spectra of
vorticity and divergence both show variance peaks at submesoscales near the wavelength of 8km (this scale
can be well resolved in the HR simulation). The vorticity at submesoscales is dramatically reduced below the
ML; while the mesoscale vorticity variance does not change much as a result of the mesoscale eddies with a
length scale of ∼100km (Figures15a, 15d, and15g). It is no doubt that submesoscale divergence variance
should be stronger in the ML than below (Figure15b). If we take R= 0.1 as the rationale for partitioning
balanced and unbalanced motions (R > 0.1 means that the unbalanced motions cannot be ignored), the
boundary line is roughly near 10km (Figures15c, 15f, and15i), which is coincidentally consistent with the
wavelength of an effective forward kinetic energy cascade (Cao etal.,2021). Note that this is not to define the
scale of submesoscale eddies but roughly identify the scale range where unbalanced motions become nonneg-
ligible. Again, the simulated IGW should be largely underestimated in the absence of tide and high-frequency
wind forcing. In effect, the inertia-gravity waves in the northwestern Pacific upper ocean were found highly
energetic (Jing & Wu,2014).
Based on the spectral analysis, the schematic diagram diagnostically summarizes the dominant mechanisms
with roughly separated spaces: quasi-geostrophic balanced motions (QBM) at low frequencies and low
wavenumbers, linear IGW (high frequency and low wavenumber), and unbalanced submesoscale motions
(USM) at high frequencies and high wavenumbers (Figure16a). With the partition for QBM, IGW, and
USM, it is possible to diagnose their effects on vertical motions by filtering. Figure16b compares the
vertical profiles of root-mean-square (RMS) vertical velocity in spaces dominated by QBM, IGW, and
USM. Here, the Coriolis frequency f and the wavelength of 10km are used to separate the spaces. The USM
vertical motions are remarkably larger than that by either IGW or QBM, reaching ∼27m/day (RMS) below
z=−100m. The USM is likely overestimated as a part of high-mode IGW are included. In the high-fre-
quency and low-wavenumber space, the IGW-induced w continuously increases with depth, demonstrating
the enhancement of IGW at depths. In contrast, the QBM exhibit the least significant contribution to vertical
exchange in the upper ocean.
Journal of Geophysical Research: Oceans
CAO AND JING
10.1029/2021JC017812
17 of 22
5. Summary
This study investigates the submesoscale ageostrophic motions within and below the ML of the northwestern
Pacific upper ocean using a ∼500-m resolution simulation. The purpose of this work is to improve our under-
standing of the enhanced vertical velocity associated with different processes at submesoscales and its impact on
vertical heat transport. The dynamical features and mechanisms of these ageostrophic motions are diagnostically
analyzed. The scale ranges of these processes are also identified using frequency-wavenumber spectrum analysis.
The results are summarized as follows.
The submesoscale activities within and below the ML display different characteristics, indicating different
dynamical mechanisms. There exhibits strong submesoscale eddies and fronts within the ML, which are clearly
weakened below the ML (Figures2 and3); while the vertical motions are intensified at depths (recall Figure14).
Care is needed to understand the submesoscale processes below the ML because the classical variables (e.g., Ro)
for evaluating submesoscale dynamics are no longer remarkable. The Richardson number is still significant in
diagnosing the instability of the flow.
Figure 14. Frequency-wavenumber spectra of vertical vorticity (left column), horizontal buoyancy gradient (middle column), and vertical velocity (right column)
multiplied by wavenumber and frequency at z=−10m, −50m, and −200m. The dispersion relation curves of linear internal gravity waves (IGW; mode 1 and 10) are
marked, and the Coriolis frequency is indicated by the dashed line.
Journal of Geophysical Research: Oceans
CAO AND JING
10.1029/2021JC017812
18 of 22
The results show that the large vertical velocity below the ML is primarily driven by the USM over the upper
200m. The potential mechanisms for the enhanced vertical velocity arise from frontogenesis, along-isopycnal
motions, and high-mode IGW. These processes are associated with frontal effects near the Kuroshio, known as
ageostrophic frontal dynamics (Siegelman,2020). Importantly, these motions can significantly contribute to the
VHF between the ML and the ocean interior. Since the high-order IGW are not exactly qualified in this study,
classifying these wave motions into the USM would lead to an overestimate of the USM vertical velocity in
Figure16b.
The smaller submesoscales are strongly ageostrophic as shown in the F-K spectra (Figure14). The ratio (R)
between unbalanced and balanced KE can be used to approximately identify the upper bound of unbalanced
motions. The vertical velocity increases with the decrease of length scales. It is interesting that the estimated
bound is close to the length scale where forward kinetic energy transfers occur (Cao etal.,2021; the energy trans-
fers were estimated using the same data). From this perspective, the parameter, R, may also be used as a metric
for the length scales of forward kinetic energy cascade induced by unbalanced submesoscale motions.
Figure 15. Frequency-wavenumber spectra of vertical vorticity (left column) and horizontal divergence (middle column) at z=−10m, −50m, and −200m. The
right column shows the ratio between horizontal divergence and vertical vorticity in the spectral space, where red, white, and blue shadings mean R>0.1, R=0.1, and
R<0.1, respectively. The dispersion relation curves of linear internal gravity waves (IGW; mode 1 and 10) are marked, and the Coriolis frequency is indicated by the
dashed line.
Journal of Geophysical Research: Oceans
CAO AND JING
10.1029/2021JC017812
19 of 22
Essentially, the Kuroshio jet provides the energy source for the mesoscale and submesoscale activities within
and below the ML. The fronts or filaments can be arrested and advected by sheared flows in the ML and below
(Siegelman,2020). Within the ML with weak stratification, the ML eddies should be a dominant mode through
ML instabilities. Below the ML, submesoscale eddies would be clearly weakened, while the lateral buoyancy
gradient could still exist as a consequence of the spatial variation of the ML base. Although the submesoscale
ageostrophic motions are not a significant reservoir of KE (Ferrari & Wunsch,2009), they significantly contrib-
ute to the vertical exchange of tracers between the ML and the oceanic interior, e.g., supplying nutrients to
the euphotic layer (Klein & Lapeyre,2009; Zhang etal., 2019). Recent observations (e.g., Zhang etal.,2021)
investigated the ageostrophic motions at frequencies less than f. However, the processes with smaller length
scales and higher frequencies (>f) shown in this study show a stronger impact on vertical communication. Quan-
tification of these effects from in situ observations remains a great challenge. In addition, the seasonal difference
of these effects should be interesting as the submesoscale processes are seasonally modulated by the ML depth,
flow shear, atmospheric forcing, and so on.
Data Availability Statement
We thank Baylor Fox-Kemper of Brown University for valuable discussions. The authors would like to thank
NASA for the QuikSCAT data, NOAA ICOADS (http://icoads.noaa.gov), and SODA (https://www2.atmos.umd.
edu/ocean/) used as the forcing (http://podaac.jpl.nasa.gov). The source data for ROMS simulations are available
at the scientific database of South China Sea Institute of Oceanology (www.scsio.csdb.cn). The model data used
in this study have been uploaded to a public online repository (https://github.com/ROMSKURO/ROMS_KURO/
releases), which has been linked to Zenodo at https://zenodo.org/record/5613405#.YXukpZpBxjU (doi:http://
doi.org/10.5281/zenodo.5613405).
References
Bachman, S. D., Taylor, J., Adams, K., & Hosegood, P. (2017). Mesoscale and submesoscale effects on mixed layer depth in the Southern Ocean.
Journal of Physical Oceanography, 47(9), 2173–2188. https://doi.org/10.17863/CAM.10896
Balwada, S. D., Smith, K. S., & Abernathey, R. (2018). Submesoscale vertical velocities enhance tracer subduction in an idealized Antarctic
circumpolar current. Geophysical Research Letters, 45, 9790–9802. https://doi.org/10.1029/2018GL079244
Barkan, R., Molemaker, M. J., Srinivasan, K., McWilliams, J. C., & D’Asaro, E. A. (2019). The role of horizontal divergence in submesoscale
frontogenesis. Journal of Physical Oceanography, 49(6), 1593–1618. https://doi.org/10.1175/JPO-D-18-0162.1
Figure 16. (a) The diagram for the scale partition in the frequency-wavenumber space: QBM for quasi-geostrophic balanced motions (blue shading); USM for
unbalanced submesoscale motions (red shading); IGW for inertia-gravity waves (dispersion lines of mode 1, 3, and 10 are marked). A wave speed line of 20m/day is
indicated. The Coriolis frequency is also marked with a dashed line. (b) Vertical profiles of root-mean-square (RMS) w for QBM, IGW, and USM using the filtering of
Coriolis frequency and the length scale of 10km.
Acknowledgments
This work was supported by the National
Key Research and Development
Program of China (2017YFA0604104),
the National Natural Science Founda-
tion of China (92058201, 42176004,
51709092, 42149907, 41776040, and
41830538). This work was also partly
by other projects (ZDBS-LY-DQC011,
ZDRW-XH-2019-2, ISEE2021PY01, and
GML2019ZD0303).
Journal of Geophysical Research: Oceans
CAO AND JING
10.1029/2021JC017812
20 of 22
Boccaletti, G., Ferrari, R., & Fox-Kemper, B. (2007). Mixed layer instabilities and restratification. Journal of Physical Oceanography, 37,
2228–2250. https://doi.org/10.1175/jpo.3101.1
Buckingham, C. E., Gula, J., & Carton, X. (2021a). The role of curvature in modifying frontal instabilities. Part I: Review of theory and presenta-
tion of a nondimensional instability criterion. Journal of Physical Oceanography, 51(2), 299–315. https://doi.org/10.1175/JPO-D-19-0265.1
Buckingham, C. E., Gula, J., & Carton, X. (2021b). The role of curvature in modifying frontal instabilities. Part II: Application of the crite-
rion to curved density fronts at low Richardson numbers. Journal of Physical Oceanography, 51(2), 317–341. https://doi.org/10.1175/
JPO-D-20-0258.1
Bühler, O., Callies, J., & Ferrari, R. (2014). Wave-vor tex decomposition of one-dimensional ship-track data. Journal of Fluid Mechanics, 756,
1007–1026. https://doi.org/10.1017/jfm.2014.488
Calil, P. H. R., & Richards, K. J. (2010). Transient upwelling hot spots in the oligotrophic North Pacific. Journal of Geophysical Research, 115,
C02003. https://doi.org/10.1029/2009JC005360
Callies, J., Flierl, G., Ferrari, R., & Fox-Kemper, B. (2016). The role of mixed-layer instabilities in submesoscale turbulence. Journal of Fluid
Mechanics, 788, 5–41. https://doi.org/10.1017/jfm.2015.700
Cao, H., Fox-Kemper, B., & Jing, Z. (2021). Submesoscale eddies in the upper ocean of the Kuroshio Extension from high-resolution Simulation:
Energy budget. Journal of Physical Oceanography, 51(7), 2181–2201. https://doi.org/10.1175/JPO-D-20-0267.1
Cao, H., Jing, Z., Fox-Kemper, B., Yan, T., & Qi, Y. (2019). Scale transition from geostrophic motions to internal waves in the northern South
China Sea. Journal of Geophysical Research: Oceans, 124, 9364–9383. https://doi.org/10.1029/2019JC015575
Capet, X., McWilliams, J., Molemaker, M., & Shchepetkin, A. (2008a). Mesoscale to submesoscale transition in the California Current system.
Part III: Energy balance and flux. Journal of Physical Oceanography, 38, 2256–2269. https://doi.org/10.1175/2008jpo3810.1
Capet, X., McWilliams, J., Molemaker, M., & Shchepetkin, A. (2008b). Mesoscale to submesoscale transition in the California Current system.
Part I: Flow structure, eddy flux, and observational tests. Journal of Physical Oceanography, 38, 29–43. https://doi.org/10.1175/2007jpo3671.1
Capet, X., Roullet, G., Klein, P., & Maze, G. (2016). Intensif ication of upper-ocean submesoscale turbulence through Charney baroclinic insta-
bility. Journal of Physical Oceanography, 46(11), 3365–3384. https://doi.org/10.1175/JPO-D-16-0050.1
Carton, J., & Giese, B. (2008). A reanalysis of ocean climate using Simple Ocean Data Assimilation (SODA). Monthly Weather Review, 136(8),
2999–3017. https://doi.org/10.1175/2007mwr1978.1
Chelton, D. B., Gaube, P., Schlax, M. G., Early, J. J., & Samelson, R. M. (2011). The influence of nonlinear mesoscale eddies on near-surface
oceanic chlorophyll. Science, 334(6054), 328–332. https://doi.org/10.1126/science.1208897
Chelton, D. B., Schlax, M. G., Samelson, R. M., & De Szoeke, R. A. (2007). Global observations of large oceanic eddies. Geophysical Research
Letters, 34, L15606. https://doi.org/10.1029/2007GL030812
D’addezio, J. M., Jacobs, G. A., & Yaremchuk, M. (2020). Submesoscale eddy vertical covariances and dynamical constraints from high-resolu-
tion numerical simulations. Journal of Physical Oceanography, 50(4), 1087–1115. https://doi.org/10.1175/JPO-D-19-0100.1
Danioux, E., Vanneste, J., Klein, P., & Sasaki, H. (2012). Spontaneous inertia-gravity-wave generation by surface-intensified turbulence. Journal
of Fluid Mechanics, 699, 153–173. https://doi.org/10.1017/jfm.2012.90
D’Asaro, E., Lee, C., Rainville, L., Harcourt, R., & Thomas, L. (2011). Enhanced turbulence and energy dissipation at ocean fronts. Science,
332(6027), 318–322. https://doi.org/10.1126/science.1201515
de Boyer Montégut, C., Madec, G., Fischer, A. S., Lazar, A., & Iudicone, D. (2004). Mixed layer depth over the global ocean: An examination of
profile data and a profile-based climatology. Journal of Geophysical Research, 109, C12003. https://doi.org/10.1029/2004JC002378
Ferrari, R., & Wunsch, C. (2009). Ocean circulation kinetic energy: Reservoirs, sources, and sinks. Annual Review of Fluid Mechanics, 41,
253–282. https://doi.org/10.1146/annurev.fluid.40.111406.102139
Fox-Kemper, B., Ferrari, R., & Hallberg, R. (2008). Parameterization of mixed layer eddies. Part I: Theory and diagnosis. Journal of Physical
Oceanography, 38, 1145–1165. https://doi.org/10.1175/2007jpo3792.1
Freilich, M. A., & Mahadevan, A. (2019). Decomposition of vertical velocity for nutrient transport in the upper ocean. Journal of Physical Ocean-
ography, 49(6), 1561–1575. https://doi.org/10.1175/JPO-D-19-0002.1
Fu, L.-L., Chelton, D. B., Le Traon, P.-Y., & Morrow, R. (2010). Eddy dynamics from satellite altimetry. Oceanography, 23(4), 14–25. https://
doi.org/10.5670/oceanog.2010.02
Gula, J., Molemaker, M., & McWilliams, J. C. (2014). Submesoscale cold filaments in the Gulf Stream. Journal of Physical Oceanography,
44(10), 2617–2643. https://doi.org/10.1175/JPO-D-14-0029.1
Hoskins, B. J., & Bretherton, F. P. (1972). Atmospheric frontogenesis models: Mathematical formulation and solution. Journal of the Atmos-
pheric Sciences, 29, 11–37. https://doi.org/10.1175/1520-0469(1972)029<0011:AFMMFA>2.0.CO;2
Hoskins, B. J., Draghici, I., & Davies, H. C. (1978). A new look at the ω-equation. Quarterly Journal of the Royal Meteorological Society, 104,
31–38. https://doi.org/10.1002/qj.49710443903
Hua, B. L., & Klein, P. (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
Huang, X., Jing, Z., Zheng, R., & Cao, H. (2020). Dynamical analysis of submesoscale fronts associated with wind-forced offshore jet in the
western South China Sea. Acta Oceanologica Sinica, 39, 1–12. https://doi.org/10.1007/s13131-020-1671-4
Jing, Z., Fox-Kemper, B., Cao, H., Zheng, R., & Du, Y. (2021). Submesoscale fronts and their dynamical processes associated with symmet-
ric instability in the Northwest Pacific subtropical ocean. Journal of Physical Oceanography, 51(1), 83–100. https://doi.org/10.1175/
JPO-D-20-0076.1
Jing, Z., & Wu, L. (2014). Intensified diapycnal mixing in the midlatitude western boundary currents. Scientific Reports, 4, 7412. https://doi.
org/10.1038/srep07412
Johnson, L., Lee, C. M., D’Asaro, E. A., Thomas, L., & Shcherbina, A. (2020a). Restratification at a California current upwelling front. Part I:
Observations. Journal of Physical Oceanography, 50, 1455–1472. https://doi.org/10.1175/JPO-D-19-0203.1
Johnson, L., Lee, C. M., D’Asaro, E. A., Wenegrat, J. O., & Thomas, L. N. (2020b). Restratification at a California current upwelling front. Part
II: Dynamics. Journal of Physical Oceanography, 50, 1473–1487. https://doi.org/10.1175/JPO-D-19-0204.1
Kaneko, H., Yasuda, I., Komatsu, K., & Itoh, S. (2013). Observations of vertical turbulent nitrate flux across the Kuroshio. Geophysical Research
Letters, 40, 3123–3127. https://doi.org/10.1002/grl.50613
Kara, A. B., Rochford, P. A., & Hurlburt, H. E. (2003). Mixed layer depth variability over the global ocean. Journal of Geophysical Research,
108(C3), 3079. https://doi.org/10.1029/2000JC000736
Klein, P., Hua, B. L., Lapeyre, G., Capet, X., Gentil, S. L., & Sasaki, H. (2008). Upper ocean turbulence from high-resolution 3D simulations.
Journal of Physical Oceanography, 38, 1748–1763. https://doi.org/10.1175/2007jpo3773.1
Klein, P., & Lapeyre, G. (2009). The oceanic vertical pump induced by mesoscale and submesoscale turbulence. Annual Review of Marine
Science, 1(1), 351–375. https://doi.org/10.1146/annurev.marine.010908.163704
Journal of Geophysical Research: Oceans
CAO AND JING
10.1029/2021JC017812
21 of 22
Klymak, J. M., Shearman, R. K., Gula, J., Lee, C. M., D’Asaro, E. A., Thomas, L. N., etal. (2016). Submesoscale streamers exchange water on
the north wall of the Gulf Stream. Geophysical Research Letters, 43, 1226–1233. https://doi.org/10.1002/2015GL067152
Lapeyre, G., Klein, P., & Hua, B. L. (2006). Oceanic restratification forced by surface frontogenesis. Journal of Physical Oceanography, 36(8),
1577–1590. https://doi.org/10.1175/JPO2923.1
Large, W., McWilliams, J. C., & Doney, S. (1994). Oceanic vertical mixing: A review and a model with a nonlocal boundary layer parameteriza-
tion. Reviews of Geophysics, 32, 363–403. https://doi.org/10.1029/94RG01872
Lazaneo, C. Z., Napolitano, D. C., Silveira, I. C. A., Tandon, A., MacDonald, D. G., Avila, R. A., etal. (2020). On the role of turbulent mixing
produced by vertical shear between the Brazil Current and the intermediate western boundary current. Journal of Geophysical Research:
Oceans, 125, e2019JC015338. https://doi.org/10.1029/2019JC015338
Lévy, M., Iovino, D., Resplandy, L., Klein, P., Madec, G., Tréguier, A., etal. (2012). Large-scale impacts of submesoscale dynamics on phyto-
plankton: Local and remote effects. Ocean Modelling, 43–44, 77–93. https://doi.org/10.1016/j.ocemod.2011.12.003
Lévy, M., Klein, P., & Tréguier, A. (2001). Impact of sub-mesoscale physics on production and subduction of phytoplankton in an oligotrophic
regime. Journal of Marine Research, 59, 535–565. https://doi.org/10.1357/002224001762842181
Liu, F., Tang, S., & Chen, C. (2015). Satellite observations of the small-scale cyclonic eddies in the western South China Sea. Biogeosciences,
12(2), 299–305. https://doi.org/10.5194/bg-12-299-2015
Liu, Z., Liao, G., Hu, X., & Zhou, B. (2020). Aspect ratio of eddies inferred from Argo floats and satellite altimeter data in the ocean. Journal of
Geophysical Research: Oceans, 125, e2019JC015555. https://doi.org/10.1029/2019JC015555
Luo, S., Jing, Z., & Qi, Y. (2020). Submesoscale flows associated with convergent strain in an anticyclonic eddy of the Kuroshio Extension: A
high-resolution numerical study. Ocean Science Journal, 55(2), 249–264. https://doi.org/10.1007/s12601-020-0022-x
Mahadevan, A., Pascual, A., Rudnick, D. L., Ruiz, S., Tintoré, J., & D’Asaro, E. (2020). Coherent pathways for vertical transport from the surface
ocean to interior. Bulletin of the American Meteorological Society, 101(11), E1996–E2004. https://doi.org/10.1175/BAMS-D-19-0305.1
Mahadevan, A., & Tandon, A. (2006). An analysis of mechanisms for submesoscale vertical motion at ocean fronts. Ocean Modelling, 14,
241–256. https://doi.org/10.1016/j.ocemod.2006.05.006
McWilliams, J. C. (2016). Submesoscale currents in the ocean. Proceedings of the Royal Society A: Mathematical, Physical and Engineering
Science, 472(2189), 20160117. https://doi.org/10.1098/rspa.2016.0117
McWilliams, J. C. (2017). Submesoscale surface fronts and filaments: Secondary circulation, buoyancy flux, and frontogenesis. Journal of Fluid
Mechanics, 823, 391–432. https://doi.org/10.1017/jfm.2017.294
McWilliams, J. C., Gula, J., Molemaker, M. J., Renault, L., & Shchepetkin, A. F. (2015). Filament frontogenesis by boundary layer turbulence.
Journal of Physical Oceanography, 45, 1988–2005. https://doi.org/10.1175/JPO-D-14-0211.1
Nagai, T., Gruber, N., Frenzel, H., Lachkar, Z., McWilliams, J. C., & Plattner, G.-K. (2015). Dominant role of eddies and filaments in the offshore
transport of carbon and nutrients in the California Current System. Journal of Geophysical Research: Oceans, 120, 5318–5341. https://doi.
org/10.1002/2015JC010889
Plougonven, R., & Snyder, C. (2007). Inertia-gravity waves spontaneously generated by jets and fronts. Part I: Different baroclinic life cycles.
Journal of the Atmospheric Sciences, 64, 2502–2520. https://doi.org/10.1175/JAS3953.1
Qiu, B., & Chen, S. (2013). Concurrent decadal mesoscale eddy modulations in the western North Pacific Subtropical Gyre. Journal of Physical
Oceanography, 43(2), 344–358.
Qiu, B., Chen, S., Klein, P., Sasaki, H., & Sasai, Y. (2014). Seasonal mesoscale and submesoscale eddy variability along the north pacific subtrop-
ical countercurrent. Journal of Physical Oceanography, 44(12), 3079–3098. https://doi.org/10.1175/JPO-D-14-0071.1
Qiu, B., Chen, S., Klein, P., Torres, H., Wang, J., Fu, L.-L., & Menemenlis, D. (2020). Reconstructing upper ocean vertical velocity field from sea
surface height in the presence of unbalanced motion. Journal of Physical Oceanography, 50, 55–79. https://doi.org/10.1175/JPO-D-19-0172.1
Qiu, B., Chen, S., Klein, P., Ubelmann, C., Fu, L.-L., & Sasaki, H. (2016). Reconstructability of three-dimensional upper ocean circulation from
SWOT sea surface height measurements. Journal of Physical Oceanography, 46, 947–963. https://doi.org/10.1175/JPO-D-15-0188.1
Rocha, C. B., Gille, S., Chereskin, T., & Menemenlis, D. (2016). Seasonality of submesoscale dynamics in the Kuroshio Extension. Geophysical
Research Letters, 43, 11304–11311. https://doi.org/10.1002/2016GL071349
Rosso, I., Hogg, A., Strutton, P., Kiss, A., Matear, R., Klocker, A., etal. (2014). Vertical transport in the ocean due to submesoscale structures:
Impacts in the Kerguelen region. Ocean Modelling, 80, 10–23. https://doi.org/10.1016/j.ocemod.2014.05.001
Rudnick, D. L. (2001). On the skewness of vorticity in the upper ocean. Geophysical Research Letters, 28, 2045–2048. https://doi.
org/10.1029/2000GL012265
Sasaki, H., Klein, P., Qiu, B., & Sasai, Y. (2014). Impact of oceanic-scale interactions on the seasonal modulation of ocean dynamics by the
atmosphere. Nature Communications, 5(1), 5636. https://doi.org/10.1038/ncomms6636
Schubert, R., Gula, J., Greatbatch, R. J., Baschek, B., & Biastoch, A. (2020). The submesoscale kinetic energy cascade: Mesoscale absorp-
tion of submesoscale mixed-layer eddies and frontal downscale fluxes. Journal of Physical Oceanography, 50(9), 2573–2589. https://doi.
org/10.1175/JPO-D-19-0311.1
Shakespeare, C., & Taylor, J. (2015). The spontaneous generation of inertia-gravity waves during frontogenesis forced by large strain: Numerical
solutions. Journal of Fluid Mechanics, 772, 508–534. https://doi.org/10.1017/jfm.2015.197
Shchepetkin, A., & McWilliams, J. C. (2005). The Regional Oceanic Modeling System (ROMS): A split-explicit, free-surface, topography-fol-
lowing-coordinate oceanic model. Ocean Modelling, 9(4), 347–404. https://doi.org/10.1016/j.ocemod.2004.08.002
Shcherbina, A., D’Asaro, E., Lee, C., Klymak, J., Molemaker, M., & McWilliams, J. C. (2013). Statistics of vertical vorticity, divergence, and
strain in a developed submesoscale turbulence field. Geophysical Research Letters, 40, 4706–4711. https://doi.org/10.1002/grl.50919
Siegelman, L. (2020). Energetic submesoscale dynamics in the ocean interior. Journal of Physical Oceanography, 50, 727–749. https://doi.
org/10.1175/JPO-D-19-0253.1
Siegelman, L., Klein, P., Rivière, P., Thompson, A. F., Torres, H. S., Flexas, M., etal. (2020). Enhanced upward heat transport at deep submesos-
cale ocean fronts. Nature Geoscience, 13(1), 50–55. https://doi.org/10.1038/s41561-019-0489-1
Srinivasan, K., McWilliams, J. C., Molemaker, M. J., & Barkan, R. (2019). Submesoscale vortical wakes in the lee of topography. Journal of
Physical Oceanography, 49(7), 1949–1971. https://doi.org/10.1175/JPO-D-18-0042.1
Su, Z., Wang, J., Klein, P., Thompson, A., & Menemenlis, D. (2018). Ocean submesoscales as a key component of the global heat budget. Nature
Communications, 9, 775. https://doi.org/10.1038/s41467-018-02983-w
Thomas, L. N., Tandon, A., & Mahadevan, A. (2008). Submesoscale processes and dynamics. In M. W. Hecht, & H. Hasumi (Eds.), Ocean mode-
ling in an eddying regime, Geophysical Monograph Series (Vol. 177, pp. 17–38). https://doi.org/10.1029/177gm04
Thomas, L. N., Taylor, J., Ferrari, R., & Joyce, T. (2013). Symmetric instability in the Gulf Stream. Deep-Sea Research Part II Topical Studies in
Oceanography, 91, 96–110. https://doi.org/10.1016/j.dsr2.2013.02.025
Journal of Geophysical Research: Oceans
CAO AND JING
10.1029/2021JC017812
22 of 22
Thomas, L. N., Taylor, J. R., D’Asaro, E. A., Lee, C. M., Klymak, J. M., & Shcherbina, A. (2016). Symmetric instability, inertial oscillations, and
turbulence at the Gulf Stream front. Journal of Physical Oceanography, 46(1), 197–217. https://doi.org/10.1175/JPO-D-15-0008.1
Thompson, A. F., Lazar, A., Buckingham, C., Naveira Garabato, A. C., Damerell, G. M., & Heywood, K. J. (2016). Open-ocean submesoscale
motions: A full seasonal cycle of mixed layer instabilities from gliders. Journal of Physical Oceanography, 46(4), 1285–1307. https://doi.
org/10.1175/JPO-D-15-0170.1
Torres, H. S., Klein, P., Menemenlis, D., Qiu, B., Su, Z., Wang, J., etal. (2018). Partitioning ocean motions into balanced motions and internal
gravity waves: A modeling study in anticipation of future space missions. Journal of Geophysical Research: Oceans, 123, 8084–8105. https://
doi.org/10.1029/2018JC014438
Vanneste, J. (2013). Balance and spontaneous wave generation in geophysical flows. Annual Review of Fluid Mechanics, 45(1), 147–172. https://
doi.org/10.1146/annurev-fluid-011212-140730
Wang, S., Jing, Z., Liu, H., & Wu, L. (2018). Spatial and seasonal variations of submesoscale eddies in the eastern tropical Pacific Ocean. Journal
of Physical Oceanography, 48(1), 101–116. https://doi.org/10.1175/JPO-D-17-0070.1
Woodruff, S., Worley, S., Lubker, S., Ji, Z., Eric Freeman, J., Berry, D., etal. (2011). ICOADS release 2.5: Extensions and enhancements to the
surface marine meteorological archive. International Journal of Climatology, 31(7), 951–967. https://doi.org/10.1002/joc.2103
Xu, L., Li, P., Xie, S., Liu, Q., Liu, C., & Gao, W. (2016). Observing mesoscale eddy effects on mode-water subduction and transport in the North
Pacific. Nature Communications, 7(1). 10505. https://doi.org/10.1038/ncomms10505
Yu, J., Zheng, Q., & Jing, Z. (2018). Satellite observations of sub-mesoscale vortex trains in the western boundary of the South China Sea. Jour-
nal of Marine Systems, 183, 56–62. https://doi.org/10.1016/j.jmarsys.2018.03.010
Yu, X., Garabato, A. C., Martin, A. P., Buckingham, C. E., Brannigan, L., & Su, Z. (2019a). An annual cycle of submesoscale vertical flow and
restratification in the upper ocean. Journal of Physical Oceanography, 49(6), 1439–1461. https://doi.org/10.1175/JPO-D-18-0253.1
Yu, X., Naveiragarabato, A. C., Martin, A. P., Evans, D. G., & Su, Z. (2019b). Wind-forced symmetric instability at a transient mid-ocean front.
Geophysical Research Letters, 46, 11281–11291. https://doi.org/10.1029/2019GL084309
Zeng, X., Belkin, I. M., Peng, S., & Li, Y. (2014). East Hainan upwelling fronts detected by remote sensing and modelled in summer. Interna-
tional Journal of Remote Sensing, 35(11–12), 4441–4451. https://doi.org/10.1080/01431161.2014.916443
Zhang, Z., Qiu, B., Klein, P., & Travis, S. (2019). The influence of geostrophic strain on oceanic ageostrophic motion and surface chlorophyll.
Nature Communications, 10, 2838. https://doi.org/10.1038/s41467-019-10883-w
Zhang, Z., Tian, J., Qiu, B., Zhao, W., Chang, P., Wu, D., etal. (2016). Observed 3D structure, generation, and dissipation of oceanic mesoscale
eddies in the South China Sea. Scientific Reports, 6, 24349. https://doi.org/10.1038/srep24349
Zhang, Z., Wang, W., & Qiu, B. (2014). Oceanic mass transport by mesoscale eddies. Science, 345(6194), 322–324. https://doi.org/10.1126/
science.1252418
Zhang, Z., Zhang, X., Qiu, B., Zhao, W., Zhou, C., Huang, X., etal. (2021). Submesoscale currents in the subtropical upper ocean observed by
two-year long high-resolution mooring arrays. Journal of Physical Oceanography, 51(1), 187–206. https://doi.org/10.1175/JPO-D-20-0100.1
Zheng, Q. (2017). Satellite SAR detection of sub-mesoscale ocean dynamic processes. (pp. 121–178). World Scientific.
... Wang et al., 2021), and the regulation of climate variability (e.g., Taylor & Thompson, 2023). Submesoscale vertical heat transport (SVHT) can be much larger than mesoscale vertical heat transport (MVHT), and comparable to or even larger than heat flux at the air-sea interface, exerting a significant impact on the climate system (Cao & Jing, 2022;Johnson et al., 2020;Su et al., 2018Su et al., , 2020. This study aims to estimate SVHT and discuss its governing mechanisms at the Kuroshio Extension (KE). ...
... where w and T denote the vertical velocity and temperature, respectively. Here the density constant ρ 0 is chosen to be 1027.5 kg m 3 and the specific heat capacity of seawater C p is 3,985 J kg 1 K 1 (Cao & Jing, 2022;Siegelman, 2020;Siegelman et al., 2020). RVHT represents the part of vertical heat transport induced by the mesoand submeso-scale coupling; it includes both the transport of submesocale temperature anomaly by mesoscale motions and the transport of mesoscale temperature anomaly by submesoscale motions. ...
... The domain-averaged SVHT, MVHT and RVHT from Equation 2 are all sensitive to the choice of filter. For example, results for the depth of 40 m are provided (see Figure S3 in Supporting Information S1 Previous studies have found that improving spatial, both horizontal and vertical, and temporal resolution would lead to enhanced submesoscale motions and therefore larger SVHT (Cao & Jing, 2022;Siegelman, 2020;Su et al., 2020). Using the MITgcm LLC4320 model simulation, we find that RVHT based on the Butterworth filter is much smaller than that based on other two filters. ...
Article
Full-text available
Submesoscale processes significantly influence the air‐sea interaction, heat budget and the distribution of physical and biogeochemical tracers in the upper‐ocean. We study the spatiotemporal characteristics and generation mechanisms of submesoscale vertical heat transport (SVHT) at the Kuroshio Extension using a submesoscale‐permitting simulation. Compared with the commonly used Boxcar and Hanning filters, the clean‐cut feature of the Butterworth filter in the wavenumber domain makes it a proper filter to diagnose SVHT. SVHT is systematically upward, peaking in winter. Through causality analysis, we find that, among the basic factors (mixed layer depth, strain rate, surface wind stress, and horizontal buoyancy gradient) from the conventional submesoscale generation mechanism scalings, mixed layer depth plays the leading role in modulating the intraseasonal variation of SVHT. Using the budget of submesoscale temperature variance and causality analysis, we find that advection also plays an important role in regulating SVHT. This work suggests that the choice of appropriate spatial filter is important for SVHT diagnosis and including advection may help improve submesoscale parameterization schemes.
... However, recent studies have suggested that submesoscale processes tend to be more effective in driving VHT, both within and below the mixed layer (Cao & Jing, 2022;Siegelman et al., 2020;Su et al., 2018Su et al., , 2020Wang et al., 2022;Yang et al., 2021;Yu et al., 2019), as a consequence of enhanced vertical flows typically ranging from 10 to 100 m day 1 . During the wintertime, submesoscale motions associated with mixed-layer instabilities (Boccaletti et al., 2007) can drive extremely strong upward VHT (Su et al., , 2020Yang et al., 2021). ...
... In these regions, subsurface frontogenesis associated with the confluence of background flows can lead to the development of ageostrophic secondary circulation in response to front intensification. This circulation pattern induces upwelling on the warmer side of the front and downwelling on the colder side (Cao & Jing, 2022;Klein et al., 2008;Ramachandran et al., 2014), ultimately resulting in a net upward VHT in the oceanic interior. Supporting this notion, Siegelman et al. (2020), using sea seal data, demonstrated that deep-reaching submesocale fronts in the Antarctic Circumpolar Current can trigger a transient upward VHT rate as high as 2000 W m 2 . ...
Article
Full-text available
Plain Language Summary Understanding the upper‐ocean heat budget is of great importance for gaining insight into how oceanic processes modulate the climate system, yet vertical heat transport (VHT) by submesoscale processes remains rarely studied using observations. Recently, scientists have identified the potential importance of submesoscale instabilities to enhance upward VHT within the mixed layer. However, the vertical pathways of heat from the ocean interior to the surface and the underlying mechanisms remain unclear, largely due to the limitations in observing such small, fast scales. To elucidate these questions, we conducted high‐resolution (a horizontal resolution of ∼0.6 km), synoptic in‐situ observations targeted at submesoscale phenomena near mesoscale eddies. Our study reveals substantial contributions of submesoscale processes to upward VHT in the stratified subsurface layer. This causes a notable imbalance in VHT by mesoscale, submesoscale, and mixing processes. These findings provide valuable insights for enhancing our understanding of heat uptake in the ocean.
... Limited by the observation method, these unsolved questions pose hindrances to our understanding of the submesoscale processes and their potential climatic and ecological effect. The present study investigates the vertical heat and salinity fluxes of oceanic submesoscale fronts through targeted field observations during the early boreal spring in the Kuroshio-Oyashio Extension (KOE) region (Fig. 1a), a key area abundant with energetic submesoscale fronts 42,43 . Besides, this region is also known to be a highly productive fishing ground, supporting a variety of commercially important marine fishes. ...
Article
Full-text available
Submesoscale fronts, with horizontal scale of 0.1–10 km, are key components of climate system by driving intense vertical transports of heat, salt and nutrients in the ocean. However, our knowledge on how large the vertical transport driven by one single submesoscale front can reach remains limited due to the lack of comprehensive field observations. Here, based on high-resolution in situ observations in the Kuroshio-Oyashio Extension region, we detect an exceptionally sharp submesoscale front. The oceanic temperature (salinity) changes sharply from 14 °C (34.55 psu) to 2 °C (32.7 psu) within 2 km across the front from south to north. Analysis reveals intense vertical velocities near the front reaching 170 m day⁻¹, along with upward heat transport up to 1.4 × 10⁻² °C m s⁻¹ and salinity transport reaching 4 × 10⁻⁴ psu m s⁻¹. The observed heat transport is much larger than the values reported in previous observations and is three times as that derived from current eddy-rich climate models, whereas the salinity transport enhances the nutrients concentration with prominent implications for marine ecosystem and fishery production. These observations highlight the vertical transport of submesoscale fronts and call for a proper representation of submesoscale processes in the next generation of climate models.
... Neglecting ageostrophic and nonhydrostatic effects could also lead to misestimation of vertical velocities (Mahadevan & Tandon, 2006). These effects of submesoscale along-isopycnal motions on vertical excursion of tracers have also been demonstrated in recent observational and modeling studies (Cao & Jing, 2022;Qu et al., 2022;Ruiz et al., 2019). ...
Article
Full-text available
Mesoscale and submesoscale processes have crucial impacts on ocean biogeochemistry, importantly enhancing the primary production in nutrient‐deficient ocean regions. Yet, the intricate biophysical interplay still holds mysteries. Using targeted high‐resolution in situ observations in the South China Sea, we reveal that isopycnal submesoscale stirring serves as the primary driver of vertical nutrient transport to sustain the dome‐shaped subsurface chlorophyll maximum (SCM) within a long‐lived cyclonic mesoscale eddy. Density surface doming at the eddy core increased light exposure for phytoplankton production, while along‐isopycnal submesoscale stirring disrupted the mesoscale coherence and drove significant vertical exchange of tracers. These physical processes play a crucial role in maintaining the elevated phytoplankton biomass in the eddy core. Our findings shed light on the universal mechanism of how mesoscale and submesoscale coupling enhances primary production in ocean cyclonic eddies, highlighting the pivotal role of submesoscale stirring in structuring marine ecosystems.
... More details about the model setup can be found in Cao et al. (2021) and Jing et al. (2021). Based on previous studies, simulations have shown good performance in characterizing the circulation, thermohaline structure, mesoscale eddies and submesoscale motions in the KOE region (Luo et al. 2020;Cao and Jing 2022;Dong et al. 2022). In this study, we will use the temperature, salinity, density, and velocities in the upper 2000 m from the model output. ...
Article
Full-text available
A variety of submesoscale coherent vortices (SCVs) in the Kuroshio Extension region have been reported by recent observational studies, and the preliminary understanding of their properties, spatial distribution and possible origins has progressively improved. However, due to relatively sparse in situ observations, the generation mechanisms of these SCVs and associated dynamic processes remain unclear. In this study, we use high-resolution model simulations to fill the gaps of the in situ observations in terms of the three-dimensional structures and life cycles of SCVs. Vortex detection and tracking algorithms are adopted and the characteristics of warm-core and cold-core SCVs are revealed. These vortices have finite Rossby numbers (0.25-0.4) and their horizontal structures can be well described by the Tayler vortex model in terms of the gradient wind balance. The vertical velocity field is characterized by a distinct dipole pattern with upwelling and downwelling cells at the vortex edge. It is very likely that both types of SCVs are generated along the eastern Japan coast through flow–topography interactions, and the Izu–Ogasawara Ridge and Hokkaido slope are found to be two important generation sites where topography friction produces extremely low potential vorticity. After leaving the boundary, SCVs can propagate over long distances and trap a water volume of ~10 ¹¹ m ³ .
Article
Full-text available
By providing valuable data that allow scientists to study various oceanographic characteristics on a global scale, remote sensing techniques have considerably advanced our understanding of ocean fronts. Ocean fronts involve the interaction of water masses with specific physical properties such as temperature, sea color, salinity, and density. In particular, ocean fronts can act as barriers, impeding the movement of water masses and leading to the convergence or divergence of nutrients and marine species. Research on ocean fronts and their impact on marine biodiversity and physical environments has recently become popular. This paper introduces ocean front research progress based on remote sensing images, including research material, methods, limitations, and possible future research directions. The latest research on spatiotemporal variation in ocean fronts has substantially enhanced our understanding of the interaction of water masses with specific physical properties in the ocean.
Article
Full-text available
Cross‐shelf penetrating fronts (CPFs) induce significant cross‐shelf exchange of water properties and nutrients, and thus are important to coastal environments. In this study, the characteristics and mechanisms of realistic large‐scale CPFs in the East China Sea in summer were investigated based on a data assimilative model. The model reproduced CPFs matched well with satellite observations. Although the cross‐shelf currents were predominantly offshore off the Zhe‐Min Coast, only three strong large‐scale CPFs occurred in the summer of 2014. The three‐dimensional structure of CPF in the model was similar with that observed in previous research. Two different mechanisms were responsible for the formation of observed CPFs. Two CPFs formed as a result of the convergence of the Taiwan Warm Current (TWC) and the Zhe‐Min Coastal Current (ZMCC), while the other one was caused by the undulation of thermocline. Heat budget analysis suggests that the undulation of thermocline was caused by horizontal and vertical advection. Sensitivity experiments suggest that southerly wind relaxation and tidal forcing are indispensable conditions for CPF formation. Tidal forcing makes the axis of the ZMCC shift offshore by ∼50 km, so that the ZMCC could impinge right against the axis of the TWC. The relaxation of the southerly winds allows the ZMCC to extend southward. Southerly wind relaxation in summer is mostly associated with tropical cyclones. Without winds and synoptic variation of the TWC, CPFs form periodically due to the strengthening of the ZMCC during neap tide period.
Preprint
Full-text available
Submesoscale flows (0.1 - 10 km) are often associated with large vertical velocities, which can have a significant impact on the transport of surface tracers, such as carbon. However, global models do not adequately account for these small-scale effects, which still require a proper parameterization. In this study, we introduced a passive tracer into the mixed layer of the northern Atlantic Ocean using a CROCO simulation with a high horizontal resolution of Δx = 800 m, aiming to investigate the seasonal submesoscale effects on vertical transport. Using surface vorticity and strain criteria, we identified regions with submesoscale fronts and quantified the associated subduction, that is the export of tracer below the mixed layer depth. The results suggest that the tracer vertical distribution and the contribution of frontal subduction can be estimated from surface strain and vorticity. Notably, we observed significant seasonal variations. In winter, the submesoscale fronts contribute up to 40% of the vertical advective transport of tracer below the mixed layer, while representing only 5% of the domain. Conversely, in summer, fronts account for less than 1% of the domain and do not contribute significantly to the transport below the mixed layer. The findings of this study contribute to a better understanding of the seasonal water subduction due to fronts in the region.
Article
Full-text available
Submesoscale currents are known to be associated with strong vertical velocities (O (10) m/day), regulating the redistributions of energy and matter balances. The northern South China Sea (SCS) is fulfilled with submesoscale motions, which might induce strong vertical heat transport (VHT). We set up a 1-km horizontal resolution Massachusetts Institute of Technology General Circulation Model (MITgcm) to study the seasonal variations in submesoscale vertical heat transport in shelf regions and open seas. Spectrum analysis shows that the spatial scale separating submesoscale and mesoscale motions are 14 and 30 km for the shelf and open regions, respectively. The submesoscale VHT in the shelf region is one order of magnitude larger than that in the open ocean. The former has the largest value in summer and winter, which might be induced by summer upwelling and winter downwelling, while the latter is strongest in winter and weakest in summer in open regions. The submesoscale VHT also appears to have intra-seasonal variations and might be attributed to the disturbances of tropical cyclones and life stages of submesoscale eddies. The submesoscale VHT is strongest in the pregeneration phase of the eddies, and the maximum VHT belt has an entrainment type at the developing and mature stages. The chlorophyll-a concentration also has the same temporal variation as the different life-stage of eddies. This study provides local VHT induced by submesoscale motions, which is expected to improve our understanding of submesoscale air–sea interactions and their biological effects.
Article
Full-text available
Oceanic mesoscale and submesoscale eddies produce pronounced vertical buoyancy flux, playing an important role in ocean restratification. This study used a 1-km ocean simulation to investigate the seasonality of the vertical eddy buoyancy flux (VEBF) in the Kuroshio Extension as well as its underlying dynamics. The simulated VEBF in the upper 200 m over the Kuroshio Extension has a pronounced seasonal cycle. The winter VEBF peaks in the mixed layer, whereas the summer VEBF has a much smaller magnitude but a more complicated vertical structure with a narrow peak in the shallow mixed layer and a broader and stronger peak in the seasonal thermocline. The baroclinic instability (BCI), frontogenesis and turbulent thermal wind (TTW) balance all contribute to the VEBF seasonal cycle. In winter, large surface heat loss and intense winds destroy stratification and enhance turbulent vertical mixing in the upper ocean. These phenomena intensify VEBF by promoting its components induced by the frontogenesis and TTW balance and by triggering mixed layer instability (MLI). In summer, strong stratification associated with suppressed turbulent vertical mixing weakens the contributions of the frontogenesis and TTW balance to VEBF and shifts the dominant BCI type from the MLI to the surface Charney and Philips-like types with greatly reduced growth rate compared with that of MLI in winter. The shallow peak of the VEBF in summer is mainly attributed to the TTW balance, whereas the BCI and frontogenesis account primarily for its deep peak.
Article
Full-text available
The submesoscale energy budget is complex and remains understood only in region-by-region analyses. Based on a series of nested numerical simulations, this study investigated the submesoscale energy budget and flux in the upper ocean of the Kuroshio Extension, including some innovations for examining submesoscale energy budgets in general. The highest-resolution simulation on a ~500 m grid resolves a variety of submesoscale instabilities allowing an energetic analysis in the submesoscale range. The frequency–wavenumber spectra of vertical vorticity variance (i.e., enstrophy) and horizontal divergence variance were used to identify the scales of submesoscale flows as distinct from those of inertia-gravity waves but dominating horizontal divergence variance. Next, the energy transfers between the background scales and the submesoscale were examined. The submesoscale kinetic and potential energy (SMKE and SMPE) were mainly contained in the mixed layer and energized through both barotropic (shear production) and baroclinic (buoyancy production) routes. Averaged over the upper 50 m of ROMS2, the baroclinic transfers amounted to approximately 75% of the sources for the SMKE (3.42 × 10 ⁻⁹ W/kg) versus the remaining 25% (1.12 × 10 ⁻⁹ W/kg) via barotropic downscale KE transfers. The KE field was greatly strengthened by energy sources through the boundary—this flux is larger than the mesoscale-to-submesoscale transfers in this region. Spectral energy production, importantly, reveals upscale KE transfers at larger submesoscales and downscale KE transfers at smaller submesoscales (i.e., a transition from inverse to forward KE cascade). This study seeks to extend our understanding of the energy cycle to the submesoscale and highlight the forward KE cascade induced by upper-ocean submesoscale activities in the research domain.
Article
Full-text available
This study investigates the submesoscale fronts and their dynamic effects on the mean flow due to frontal instabilities in the wind-driven summer offshore jet of the western South China Sea (WSCS), using satellite observations, a 500 m-resolution numerical simulation, and diagnostic analysis. Both satellite measurements and simulation results show that the submesoscale fronts occupying a typical lateral scale of O(∼10) km are characterized with one order of Rossby (Ro) and Richardson (Ri) numbers in the WSCS. This result implies that both geostrophic and ageostrophic motions feature in these submesoscale fronts. The diagnostic results indicate that a net cross-frontal Ekman transport driven by down-front wind forcing effectively advects cold water over warm water. By this way, the weakened local stratification and strong lateral buoyancy gradients are conducive to a negative Ertel potential vorticity (PV) and triggering frontal symmetric instability (SI) at the submesoscale density front. The cross-front ageostrophic secondary circulation caused by frontal instabilities is found to drive an enhanced vertical velocity reaching O(100) m/d. Additionally, the estimate of the down-front wind forcing the Ekman buoyancy flux (EBF) is found to be scaled with the geostrophic shear production (GSP) and buoyancy flux (BFLUX), which are the two primary energy sources for submesoscale turbulence. The large values of GSP and BFLUX at the fronts suggest an efficient downscale energy transfer from larger-scale geostrophic flows to the submesoscale turbulence owing to down-front wind forcing and frontal instabilities. In this content, submesoscale fronts and their instabilities substantially enhance the local vertical exchanges and geostrophic energy cascade towards smaller-scale. These active submesoscale processes associated density fronts and filaments likely provide new physical interpretations for the filamentary high chlorophyll concentration and frontal downscale energy transfer in the WSCS.
Article
Full-text available
Submesoscale density fronts and the associated processes of frontogenesis and symmetric instability (SI) are investigated in the Northwest Pacific subtropical counter-current (STCC) system by a high-resolution simulation and diagnostic analysis. Both satellite observations and realistic simulation show active surface fronts with a horizontal scale of ~20 km in the STCC upper ocean. Frontogenesis-induced buoyancy advection is detected to rapidly sharpen these density fronts. The direct straining effect of larger-scale geostrophic flows is a primary influence on the buoyancy-gradient frontogenetic tendency and frontal baroclinic potential vorticity (PV) enhancement. The enhanced lateral buoyancy gradients in conjunction with atmospheric forced surface buoyancy loss can produce a negative Ertel PV and trigger frontal SI in the STCC region. Up to 30% of the mixed layer (ML) inside a typical eddy has negative PV in the high-resolution simulation. As a result, the cross-front ageostrophic secondary circulations tend to restratify the surface boundary layer and induce a large vertical velocity reaching ~100 m day ⁻¹ , substantially facilitating the vertical communication of the STCC system. At the same time, the SI is identified to be responsible for a forward cascade of geostrophic kinetic energy in the STCC region, despite the coexistence of ML eddies and SI in the deep winter ML. Therefore, these active density fronts and their SI-associated submesoscale processes play important roles in the enhanced vertical exchanges (e.g., heat, nutrients and carbon) and energy transfer to smaller scales in the eddy-active STCC upper ocean, as well as triggering phytoplankton blooms at the periphery of eddies.
Article
Full-text available
We continue our study of the role of curvature in modifying frontal stability. In Part I, we obtained an instability criterion valid for curved fronts and vortices in gradient wind balance (GWB): L'q' < 0, where L' and q' are the nondimensional absolute angular momentum and Ertel potential vorticity (PV), respectively. In Part II, we investigate this criterion in a parameter space representative of low-Richardson-number fronts and vortices in GWB. An interesting outcome is that, for Richardson numbers near 1, anticyclonic flows increase in q' , while cyclonic flows decrease in q' , tending to stabilize anticyclonic and destabilize cyclonic flow. Although stability is marginal or weak for anticyclonic flow (owing to multiplication by L' ), the destabilization of cyclonic flow is pronounced, and may help to explain an observed asymmetry in the distribution of small-scale, coherent vortices in the ocean interior. We are referring to midlatitude submesoscale and polar mesoscale vortices that are generated by friction and/or buoyancy forcing within boundary layers but that are often documented outside these layers. A comparison is made between several documented vortices and predicted stability maps, providing support for the proposed mechanism. A simple expression, which is a root of the stability discriminant F' , explains the observed asymmetry in the distribution of vorticity. In conclusion, the generalized criterion is consistent with theory, observations, and recent modeling studies and demonstrates that curvature in low-stratified environments can destabilize cyclonic and stabilize anticyclonic fronts and vortices to symmetric instability. The results may have implications for Earth system models.
Article
Full-text available
In this study, we examine the role of curvature in modifying frontal stability. We first evaluate the classical criterion that the Coriolis parameter f multiplied by the Ertel potential vorticity (PV) q is positive for stable flow and that instability is possible when this quantity is negative. The first portion of this statement can be deduced from Ertel’s PV theorem, assuming an initially positive fq. Moreover, the full statement is implicit in the governing equation for the mean geostrophic flow, as the discriminant, fq, changes sign. However, for curved fronts in cyclogeostrophic or gradient wind balance (GWB), an additional term enters the discriminant owing to conservation of absolute angular momentum L. The resulting expression, (1 + Cu)fq < 0 or Lq <, 0, where Cu is a nondimensional number quantifying the curvature of the flow, simultaneously generalizes Rayleigh’s criterion by accounting for baroclinicity and Hoskins’s criterion by accounting for centrifugal effects. In particular, changes in the front’s vertical shear and stratification owing to curvature tilt the absolute vorticity vector away from its thermal wind state; in an effort to conserve the product of absolute angular momentum and Ertel PV, this modifies gradient Rossby and Richardson numbers permitted for stable flow. This forms the basis of a nondimensional expression that is valid for inviscid, curved fronts on the f plane, which can be used to classify frontal instabilities. In conclusion, the classical criterion fq , 0 should be replaced by the more general criterion for studies involving gravitational, centrifugal, and symmetric instabilities at curved density fronts. In Part II of the study, we examine interesting outcomes of the criterion applied to low-Richardson-number fronts and vortices in GWB.
Article
Full-text available
Although observational efforts have been made to detect submesoscale currents (submesoscales) in regions with deep-mixed layers and/or strong mesoscale kinetic energy (KE), there have been no long-term submesoscale observations in subtropical gyres, which are characterized by moderate values of both mixed-layer depths and mesoscale KE. In order to explore submesoscale dynamics in this oceanic regime, two nested mesoscale- and submesoscale-resolving mooring arrays were deployed in the northwestern Pacific subtropical countercurrent region during 2017–2019. Based on the two years of data, submesoscales featured by order one Rossby numbers, large vertical velocities (with magnitude of 10–50 m/day) and vertical heat flux, and strong ageostrophic KE are revealed in the upper 150 m. Although most of the submesoscales are surface intensified, they are found to penetrate far beneath the mixed layer. They are most energetic during strong mesoscale strain periods in the winter-spring season but are generally weak in the summer-autumn season. Energetics analysis suggests that the submesoscales receive KE from potential energy release but lose a portion of it through inverse cascade. Because this KE sink is smaller than the source term, a forward cascade must occur to balance the submesocale KE budget, for which symmetric instability may be a candidate mechanism. By synthesizing observations and theories, we argue that the submesoscales are generated through a combination of baroclinic instability in the upper mixed and transitional layers and mesoscale strain-induced frontogenesis, among which the former should play a more dominant role in their final generation stage.
Article
An idealized framework of steady barotropic flow past an isolated seamount in a background of constant stratification (with frequency N) and rotation (with Coriolis parameter f) is used to examine the formation, separation, instability of the turbulent bottom boundary layers (BBLs), and ultimately, the genesis of submesoscale coherent vortices (SCVs) in the ocean interior. The BBLs generate vertical vorticity ζ and potential vorticity q on slopes; the flow separates and spawns shear layers; barotropic and centrifugal shear instabilities form submesoscale vortical filaments and induce a high rate of local energy dissipation; the filaments organize into vortices that then horizontally merge and vertically align to form SCVs. These SCVs have O(1) Rossby numbers () and horizontal and vertical scales that are much larger than those of the separated shear layers and associated vortical filaments. Although the upstream flow is barotropic, downstream baroclinicity manifests in the wake, depending on the value of the nondimensional height , which is the ratio of the seamount height to that of the Taylor height , where L is the seamount half-width. When , SCVs span the vertical extent of the seamount itself. However, for , there is greater range of variation in the sizes of the SCVs in the wake, reflecting the wake baroclinicity caused by the topographic interaction. The aspect ratio of the wake SCVs has the scaling , instead of the quasigeostrophic scaling .
Article
Based on fine-resolution simulations, we present a case study of the frontogenetic generation of submesoscale flows associated with a convergence field inside an anticyclonic mesoscale eddy in the Kuroshio Extension. The results reveal that the generation of submesoscale flows on the edge of the eddy is closely related to the convergent zone in the ageostrophic secondary circulation (ASC) that develops from frontal sharpening. Diagnostic analysis of the frontal tendency based on flow decomposition shows that frontal sharpening is initiated by mesoscale strain and accelerated by submesoscale convergence. With convergence on the cold side and divergence on the warm side of fronts, submesoscale divergent fields contribute to the asymmetry of the frontal tendency because the fronts tend to be strengthened by convergent motions and weakened by divergent motions. During frontal sharpening, the cross-front ASCs generate additional positive vorticity in their downwardbranches characterized by convergence via vortex stretching, which involves neighboring submesoscale vortices. Moreover, the enhanced lateral buoyancy gradients and surface buoyancy loss are conducive to triggering symmetric instability (SI) and ageostrophic anticyclonic instability (AAI) on the edge of the eddy. In response to these submesoscale instabilities and the subsequent slumping of fronts, the removal of kinetic energy (KE) associated with shear flows is comparable to the source of submesoscale KE from the eddy, which indicates that mesoscale energy can be drained by submesoscale instabilities during frontal sharpening. This study focuses on and describes the frontogenetic generation of submesoscale flows in mesoscale eddies in the Kuroshio Extension.