1. Introduction
When ocean currents interact with topography, the energy of large-scale flow cascades to smaller scales to enhance turbulent dissipation of balanced, tidal, and internal-wave energy, as well as mixing of water properties. These interactions include critical reflection (Eriksen 1982, 1998) and scattering of internal waves (Müller and Xu 1992), internal tide generation (e.g., Garrett and Kunze 2007; Klymak et al. 2008), mean-flow generation of lee waves (e.g., Bell 1975a,b; Naveira Garabato et al. 2004; Nikurashin and Ferrari 2010; Waterman et al. 2013; Legg 2021), hydraulically controlled flow (Farmer and Armi 1999; Klymak and Gregg 2004), and vortex wakes (e.g., Baines 1995; Chang et al. 2013; MacKinnon et al. 2019; Nagai et al. 2021) and so may be important sinks for ocean energy budgets (e.g., Wunsch and Ferrari 2004).
Flow–topography interaction physics varies with flow speed U and frequency ω, stratification N, Earth’s rotation f, and topographic horizontal L and vertical H length scales (Baines 1995; Garrett and Kunze 2007). If topography has small dynamical aspect ratios NH/(fL), interaction with geostrophic and tidal currents can be described analytically with linear lee-wave or internal tide generation theories (Bell 1975a,b). Interactions become nonlinear as topography steepens (e.g., Kunze and Toole 1997; St. Laurent et al. 2003; Nikurashin et al. 2014). Stratified flow is blocked or split around tall topography (Baines 1995), exciting a submesoscale vortex wake (e.g., Chang et al. 2013; Caldeira et al. 2014; Perfect et al. 2018, 2020a; MacKinnon et al. 2019; Johnston et al. 2019) as well as lee waves (e.g., Nikurashin et al. 2014; Johnston et al. 2019; Perfect et al. 2020b). Waves and eddies of large amplitude and wavenumber will induce hydraulic jumps and instabilities, resulting in enhanced turbulent dissipation and mixing (e.g., Thorpe 1992; Polzin et al. 1997; Kunze and Toole 1997; Toole et al. 1997; Moum and Nash 2000; Kunze et al. 2002; Nash et al. 2004; Klymak et al. 2008; Kunze et al. 2012).
The Kuroshio, which is the major western boundary current of the wind-driven North Pacific subtropical gyre, flows through regions of rapidly changing complex topography. Previous microstructure measurements in Tokara Strait, where the Kuroshio interacts with islands and seamounts, have found turbulent kinetic energy (TKE) dissipation rates ε exceeding 10−7 W kg−1 and diapycnal diffusivities K exceeding 10−4 m2 s−1 that extend at least 100 km downstream of seamounts (Tsutsumi et al. 2017; Nagai et al. 2017, 2019, 2021; Hasegawa et al. 2021). Some of these elevated dissipation rates were associated with
-
high-vertical-wavenumber shear along isopycnals suggestive of near-inertial waves (Nagai et al. 2019, 2021),
-
negative potential vorticity, suggesting inertial-symmetric instabilities (Nagai et al. 2021),
-
Kelvin–Helmholtz (KH) instabilities (Hasegawa et al. 2021), and
-
bottom-boundary layer separation (Hasegawa et al. 2021),
all presumably resulting from flow–topography interactions. Downstream energetic turbulence will induce vertical mixing and inject subsurface nutrient-rich waters (Guo et al. 2012, 2013) into the sunlit surface layer (Nagai et al. 2019; Hasegawa et al. 2021) which will impact phytoplankton growth and CO2 uptake (Takahashi et al. 2009).
Strong Kuroshio–topography interactions have also been observed near Green Island, Taiwan, where TKE dissipation rates are estimated to be
Large topographic dynamical aspect ratios NH/(fL) at seamounts in Tokara Strait suggest nonlinear flow–topography interactions where geostrophic and tidal currents generate internal waves and submesoscale vortices with overlapping time and length scales (MacKinnon et al. 2019; Johnston et al. 2019; Perfect et al. 2020a,b; Puthan et al. 2021). These fluctuations can interact with each other to become nonlinear and unstable to turbulent production and dissipation. Nonlinearities complicate the interpretation of finescale flow fields, and fine- and microscale cannot be reproduced simultaneously in simulations. Thus, our understanding of the physical processes generating energetic turbulence at seamounts, that is, processes connecting finescale and microscale, is still uncertain.
Microstructure measurements of turbulence are generally time-consuming, so various parameterizations have been proposed to quantify the variability of turbulent mixing using more readily available finescale variables. Among them, the finescale parameterization (e.g., Gregg 1989; Polzin et al. 1995) has been widely validated in the interior ocean (e.g., Hibiya et al. 2012; Whalen et al. 2015) and applied (e.g., Whalen et al. 2012; Kunze 2017). However, the finescale parameterization is not expected to be valid where the dominant turbulent production physics is distinct from weakly nonlinear wave–wave interactions (McComas and Müller 1981; Henyey et al. 1986), such as bottom boundary layers (Carter and Gregg 2002; MacKinnon and Gregg 2005; Polzin et al. 2014a,b), wave–mean flow interactions (Waterman et al. 2014; Kunze and Lien 2019; Wu et al. 2023), and direct breaking of internal tides near topography (Klymak et al. 2008). In energetic internal waves, low modes can become unstable and generate turbulence (D’Asaro and Lien 2000b; MacKinnon and Gregg 2005) so that the large-eddy parameterization (Moum 1996) may be more appropriate than the finescale parameterization. Shear-driven turbulence can also be estimated using the reduced-shear parameterization (Kunze et al. 1990) if unstable shear layers with gradient Richardson numbers less than 0.25 are resolved.
In this study, electromagnetic autonomous profiling explorer (EM-APEX) profiling float and acoustic Doppler current profiler (ADCP) mooring measurements in the vicinity of Hirase Seamount in Tokara Strait are used to address two scientific questions:
-
How are velocity and density fine structure modified by flow–topography interactions?
-
Which turbulence parameterizations are appropriate in the strongly nonlinear flow field?
The remainder of this paper is organized as follows. An overview of the measurements and data analysis methods is provided in section 2. In section 3, the dynamical flow–topography regime is discussed based on nondimensional parameters calculated from the bathymetry of Hirase, background flow, and stratification. In section 4, characteristics of the finescale horizontal velocity and density fields around Hirase are described. In section 5, strong turbulent layers extending for
2. Data and methods
During November 2019, fine- and microstructure measurements were collected in Tokara Strait (30°N, Coriolis frequency f = 7.3 × 10−5 s−1) to study interactions of the Kuroshio with Hirase Seamount (Fig. 1). Hirase is a ∼400-m-tall, ∼10-km-diameter flat-top bank in ∼600-m-deep waters. The Kuroshio flowed southeastward with the surface speed of ∼1 m s−1 throughout the measurement period.
(a) Bathymetry in the East China Sea and western North Pacific overlain with monthly mean horizontal current vectors at 5-m depth during November 2019 reveal the Kuroshio passing through the measurement site (small red box) in Tokara Strait. (b) Measurement site showing Hirase (HR), Kuchinoshima (KS), and Nakanoshima (NS). Magenta dots mark upstream (MU) and downstream (MD) mooring locations, blue dots are EM-APEX float deployment points, red asterisks are float descents and surfacings, and black lines are float trajectories for the western and eastern line-array deployments. Bathymetry is from ETOPO1, and horizontal currents in (a) are from the Copernicus Marine Environment Monitoring Service (CMEMS) Global Ocean Physics Reanalysis data.
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-22-0242.1
Two moorings, each equipped with an upward-looking 75-kHz Teledyne RDI Long Ranger ADCP, were deployed upstream (MU) and downstream (MD) of the seamount (Fig. 1b). The moorings were deployed on 16–17 November and recovered on 21 November. Horizontal velocities (u, υ) were measured every 30 s in 4-m vertical bins over depths of 50–550 m. The original horizontal velocity time series are smoothed over 10 min to reduce noise. Poor-quality velocity data with large velocity errors (>0.1 m s−1) or low “Percent Good 4” values (PG4 < 90; a measure of the percentage of good-quality velocity data acquired with four ADCP beams) are also removed.
EM-APEX profiling floats were deployed in 10-float along-streamline arrays upstream of the seamount, one array on 17 November 2019 with trajectories west of Hirase and a second on 20 November 2019 with trajectories east of Hirase (Fig. 1b). Each line array was in the water for roughly 24 h while it was advected downstream of the seamount by the Kuroshio before recovery. Positions were tracked by GPS when the floats were at the surface. The float line arrays had horizontal resolution of
Each float was instrumented with two pairs of electromagnetic (EM) velocity sensors, dual fast-response FP07 thermistors, and a SeaBird Electronics SBE41 CTD (Sanford et al. 2005; Lien et al. 2016). The floats were programmed to profile continuously at a vertical speed of ∼0.15 m s−1 between the sea surface and bottom while measuring temperature T, salinity S, pressure P, and horizontal velocity (u, υ). Microscale thermal-variance dissipation rates χ were only measured during ascent.
a. Horizontal velocity
Float EM velocity sensors measure voltage differences across the instrument due to the electric field induced by conducting seawater moving in Earth’s magnetic field (Sanford et al. 1978, 2005). Measured voltages are fit to a sinusoid over half-overlapping 50-s segments to estimate horizontal velocity in ∼3.75-m depth intervals, producing independent horizontal velocities every ∼7.5 m (Lien and Sanford 2019). Horizontal velocity (u, υ) typically has less than 0.01 m s−1 uncertainty, and is relative to an unknown depth-independent constant which can vary with time and location. To obtain absolute horizontal velocity, the depth-independent constant is inferred by subtracting time-depth averaged float relative velocity from the velocity inferred from float GPS positions (Lien and Sanford 2019).
b. Vertical velocity
c. Density
Float CTD measurements are taken every 15–25 s, i.e., 2.25–3.75 m vertical resolution, and gridded to 4 m. Absolute Salinity SA, Conservative Temperature Θ, and potential density σ are computed using the Gibbs SeaWater Oceanographic Toolbox of TEOS-10 (McDougall and Barker 2011). Local stratification is calculated using Thorpe-sorted potential density σ0 referenced to 0 dbar, while unsorted σ0 is used only for computing reduced shear (sections 5 and 7c) to account for density overturns. Background potential density
d. Turbulence
FP07 microthermistors mounted on the float caps sample microscale temperature fluctuations at 125 Hz. Temperature frequency spectra ΦT(ω) are calculated in ∼5-s intervals, corresponding to ∼0.75-m depth bins. Frequency spectra are converted into vertical wavenumber spectra for temperature gradient
Measured microscale temperature could be affected by vibration and rotation of the instrument because float buoyancy is adjusted discontinuously and the float is not symmetric. In this study, deviations of observed temperature-gradient spectrum from the Kraichnan (1968) universal spectrum are quantified, and only reasonable data are used (appendix B). Float-inferred profiles of ε were consistent with concurrent VMP measurements (Nagai et al. 2019).
3. Dynamic regime of the flow–topography interaction
Qualitative properties of flow–topography interactions can be described using nondimensional parameters, i.e., lee-wave topographic Froude number Frt = mleeh0 ≈ N0h0/U, Rossby number Ro = U/(fL0), tidal excursion Ex = U/(ωtL0), and internal-tide steepness γ = mtideh0 (Garrett and Kunze 2007; Legg 2021), where mlee and mtide are lee-wave and internal tide vertical wavenumbers, h0 and L0 are vertical and horizontal scales of topography, U is the background mean or tidal current speed, N0 is background buoyancy frequency, f is Coriolis frequency, and ωtide is tidal frequency.
Harmonic analysis of the moored ADCP velocity (u, υ) profile time series in terms of a time-mean, inertial/diurnal (K1), and semidiurnal (M2) fluctuations was conducted at each depth (e.g., Codiga 2011) (Fig. 2). The upstream current is dominated by the Kuroshio, which decreases linearly with depth. At the 200-m depth of the seamount summit, U0 ≈ 0.5 m s−1 (black curve in Fig. 2). The semidiurnal tidal current Ut ≈ 0.2 m s−1 is quasi-barotropic so becomes a potentially significant contributor below 400-m depth (red dashed line in Fig. 2).
Profiles of time-mean (black), semidiurnal (M2; red), and inertial/diurnal (K1; blue) horizontal speeds
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-22-0242.1
Background stratification profiles are similar along the eastern and western trajectories (Fig. 3) with a 100-m thick late-autumn surface mixed layer, a pycnocline with
(a) Example potential density σ0 (referenced to 0 dbar) profiles near Hirase Seamount summit in the western (darker blue) and eastern (darker red) float trajectories along with background (BG) profiles (lighter colors). (b) Background stratification
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-22-0242.1
a. Kuroshio–topography interactions
At Hirase Seamount, topographic Froude number Frt ≫ 1 so the Kuroshio is energetically unable to move fluid parcels over the full height of topography, resulting in upstream flow-splitting for depths a vertical scale he = U0/N0 = 50 m below the summit (e.g., Baines 1995). Conventional linear lee-wave generation is confined to the summit crown (Nikurashin et al. 2014; Perfect et al. 2020b).
Up- and downgoing slope lee waves may be generated by topographic roughness on the flanks as flow goes around the seamount (Thorpe 1992), although this can be confounded by (i) an arrested Ekman layer shutting off bottom flow on the western flank, and (ii) formation of a thick well-mixed bottom boundary layer on the eastern flank (Trowbridge and Lentz 1991; MacCready and Rhines 1993). Lee waves will be excited by topographic features with horizontal wavenumbers f/U0 < k < N/U0 (0.1–20-km horizontal wavelengths for U0 = 0.5 m s−1 summit flow). The Rossby number Ro ∼ 1, which is analogous to the normalized lee-wave frequency |kU|/f ∼ 1, suggests generation primarily of near-inertial lee waves.
Flow-splitting around a seamount produces vortex-shedding in the downstream wake (e.g., Baines 1995). Characteristics of the wake depend on topographic Froude number Frt and Rossby number Ro measures of the strengths of stratification and rotation, respectively (Perfect et al. 2018, 2020a; Srinivasan et al. 2019, 2021). Nonlinear and highly layered wake structures are expected in the lee of Hirase Seamount for Ro ∼ 1 (Srinivasan et al. 2021). Vortices are also expected to be shed from unstable frictional boundary layers on the seamount flanks (D’Asaro 1988; Molemaker et al. 2015).
b. Tide–topography interactions
The tidal-excursion parameter Ex indicates the strength of advection or nonlinearity of tide–topography interactions (e.g., Garrett and Kunze 2007). At Hirase, Ex ≪ 1 suggests internal tide generation will be at the fundamental frequency rather than higher harmonics or quasi-steady lee waves (Bell 1975a; Mohri et al. 2010).
Steepness γ represents the ratio of bottom topographic slope to internal-tide characteristic slope (e.g., Garrett and Kunze 2007). For the semidiurnal frequency, γ > 1, so internal tides can radiate upward or downward from the summit and flanks, and beams with convective and shear instability can be excited at slope transitions (Balmforth et al. 2002; St. Laurent et al. 2003; Althaus et al. 2003; Nash et al. 2006).
The superposition of steady and oscillatory flow forcing may modulate Kuroshio–topography interactions, but this is not yet well understood. Large-eddy simulations (Puthan et al. 2021) and field observations (Chang et al. 2019; MacKinnon et al. 2019) suggest that the shedding period of wake vortices can become synchronized to subharmonics of the tidal period, and that wake velocity structures vary on the tidal cycle. However, the parameter space corresponding to Hirase conditions has not been fully explored until recently. Three-dimensional numerical simulations of flow–topography interactions at Hirase Seamount suggest that deep-reaching tidal currents can play an essential role in the deeper-layer fluctuations via (i) tidal vortex-shedding and (ii) tidal modulation of the deep current (Inoue et al. 2024).
4. Characteristics of velocity and density fields
EM-APEX float profile time series of velocity reveal that the Kuroshio is blocked or diverted by the seamount in both western and eastern trajectories with Kuroshio flow speeds ∼1 m s−1 confined above 200-m depth downstream (Figs. 4a–c). At greater depths, along- and across-stream velocities are layered downstream of the seamount, which seems to extend along with isopycnals (Figs. 4a,b,d–f). These flow characteristics are consistent with nonlinear wake structures expected for Hirase’s topographic Froude number Frt > 1 and Rossby number Ro > 1 (Srinivasan et al. 2021).
Profile time series/meridional sections of (a),(b) along-stream velocity, (d),(e) across-stream velocity, and (g),(h) local stratification along (left) western and (center) eastern EM-APEX float trajectories. The along-stream direction is defined for each profile as the average direction of the absolute current velocity in the upper 100 m, representing the Kuroshio. In (g) and (h), 4-m density overturns are blue. Charcoal shading and thick black lines follow the bathymetry at the center of each trajectory from ETOPO1; these do not always agree with the maximum depth of float measurements. Thin black contours show smoothed isopycnals. Black asterisks along the upper axes mark profiles and observation times are indicated above the upper axes of (a) and (b). Magenta labels are upstream (MU) and downstream (MD) mooring locations in (d) and (e), respectively. (right) Ensemble-average vertical profiles for (c) along-stream velocity, (f) across-stream velocity, and (i) local stratification along western (blue) and eastern trajectories (red).
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-22-0242.1
Frequency spectra for velocity and shear calculated from ADCP moorings are an order of magnitude higher than the canonical Garrett and Munk (1979) model spectrum (Fig. 5). At both up- and downstream mooring sites, frequency spectra of WKB-normalized horizontal velocity exhibit peaks at inertial/diurnal and semidiurnal frequencies (Figs. 5a,b). Inertial/diurnal signals are stronger downstream than upstream at 100–300-m depth in the pycnocline (Fig. 5a). Semidiurnal signals are stronger upstream than downstream at 300–500-m depth in the deeper layer (Fig. 5b).
Frequency spectra of WKB-normalized (a),(b) horizontal velocity and (c),(d) vertical shear obtained from upstream (black curves) and downstream (red curves) moorings along with Garrett–Munk model spectra (dashed curves). Spectra are averaged over (left) the pycnocline (100–300 m) and (right) the deeper layer (300–500 m). Dashed vertical lines indicate inertial/diurnal (D), semidiurnal (SD) and buoyancy (N) frequencies. Error bars indicate the 95% confidence intervals of spectra (Thomson and Emery 2014).
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-22-0242.1
Frequency spectra of shear are whiter and smoother than velocity spectra (Figs. 5c,d). In the pycnocline, the shear spectrum is higher downstream than upstream for frequencies ranging from subinertial to supertidal (Fig. 5c). In the deep layer where the Kuroshio is weak, up- and downstream shear spectra are almost identical at all frequencies (Fig. 5d).
Unstable stratification N2 < 0 (>4-m density overturns) is captured by floats (Figs. 4g,h), particularly in the immediate lee of Hirase in the western trajectory. Potential density profiles near the seamount summit have pronounced
5. Layers of enhanced turbulent mixing extended downstream
Around Hirase, background TKE dissipation rates ε ∼ 10−8 W kg−1 and diapycnal diffusivities K ∼ 10−4 m2 s−1 are an order of magnitude higher than in typical open-ocean pycnoclines (Fig. 6). Turbulence is particularly enhanced with ε ∼ 10−6 W kg−1 and K ∼ 10−2 m2 s−1 over the seamount flanks and at ∼150–250-m depths extending at least 20 km downstream. In the eastern deployment, a strong turbulence layer with ε ∼ 10−7 W kg−1 and K ∼ 10−2 m2 s−1 is also evident at ∼350–450-m depth, extending about 10 km downstream (Figs. 6d,f).
Depth–latitude sections (time series) of (a),(b) thermal-variance dissipation rate χ, (d),(e) turbulent kinetic energy dissipation rate ε, and (g),(h) diapycnal diffusivity K along the (left) western and (center) eastern EM-APEX float trajectories. Charcoal shading and thick black lines follow the bathymetry at the center of each trajectory from ETOPO1; these do not always agree with the maximum depth of float measurements. Thin black contours show smoothed isopycnals. Black asterisks along the upper axes mark the profiles and observation times are indicated above the upper axes of (a) and (b). Magenta labels are upstream (MU) and downstream (MD) mooring locations in (d) and (e), respectively. (right) Ensemble-average vertical profiles for (c) χ, (f) ε, and (i) K along western (blue) and eastern (red) trajectories are vertically smoothed over 40 m.
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-22-0242.1
Previous observations and numerical simulations suggest that Kuroshio–Hirase interactions shed asymmetric vortices in the downstream wake (Nagai et al. 2021; Inoue et al. 2024). These will have positive vorticity on the western flank so more likely to be stable, and negative vorticity on the eastern flank so more unstable, consistent with the observed turbulence that seems more enhanced in the eastern trajectory than western in the upper 200 m.
Enhanced TKE dissipation rates are associated with 4-m reduced shear squared |vz|2 − 4N2 > 0, suggestive of shear instability (Fig. 7). The 4-m squared gradient Froude number Fr2 = |vz|2/N2 is elevated compared to shear variance normalized by background stratification,
Relationship between turbulent kinetic energy dissipation rate ε and 4-m reduced shear |vz|2 − 4N2, where N2 is local stratification and |vz|2 vertical shear squared. Dissipation rates are binned by values of reduced shear with mean (dots) and standard error (vertical bars) of ε calculated for each bin. For clarity, bins with less than 10 data points are excluded. The 2 × 10−7 W kg−1 level (dashed horizontal line) below |vz|2 − 4N2 = 0 (dashed vertical line) represents the expected noise level for 4-m reduced shear at N ∼ 10−2 s−1.
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-22-0242.1
Vertical cross sections of (a),(b) 4-m squared gradient Froude number Fr2 = |vz|2/N2 and (d),(e) shear variance normalized by background stratification
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-22-0242.1
Mooring time series of normalized shear variance
(left) Time series of buoyancy-normalized shear variance
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-22-0242.1
The energetic turbulent layer at 150–250-m depth (Fig. 6) is consistent with previous observations of a turbulent layer extending 100 km downstream of Hirase (Nagai et al. 2017, 2019, 2021). Since isotropic turbulence is thought to persist for
Such continuous turbulence production could arise from local nonlinear instabilities of a marginally stable wake. Nagai et al. (2021) reported the turbulent layers were associated with negative potential vorticity, suggesting that inertial-symmetric or secondary Kelvin–Helmholtz instabilities play a role in far-field turbulent mixing. Unfortunately, our EM float data do not allow computation of PV. Alternatively, Kunze (2019) argued that anisotropic stratified turbulence generated at the seamount could continuously feed energy into isotropic turbulence for duration of
The turbulence layer in the pycnocline might also be caused by breaking lee waves advected downstream by the Kuroshio (Zheng and Nikurashin 2019). Because of the high topographic Froude number, lee-wave generation is confined to the seamount’s summit. These waves will be trapped within the Kuroshio by the |U| = |f/k| isotach (Kunze and Lien 2019).
Ray-tracing calculations (Lighthill 1978) are performed for summit-generated lee waves (appendix C) using observed horizontal velocity U(z) and buoyancy frequency N(z) profiles (Fig. 10b). Upward-propagating lee waves reflect either from the surface (Fig. 10a) or vertical turning points zN where |kU(zN)| = N(zN) at the base of the mixed layer. Reflected or bottom-generated downward-propagating lee waves encounter vertical critical layers at depths zf such that |kU(zf)| = f where trapping, stalling, and amplification are expected to lead to breaking and turbulence production (Kunze 1985; Kunze et al. 1995). The predicted 200–300-m critical-layer depths (Fig. 10a) bracket the observed layer of high turbulent dissipation (Fig. 6). Lee waves carried downstream in Fig. 10a have initial intrinsic frequencies |kU| < 2.5f and final vertical wavenumbers |m| = 0.02 cpm. They are carried 20 km downstream in less than 2 days. These near-inertial lee waves could be the cause of high-vertical-wavenumber shear along isopycnals in the downstream of Hirase reported by Nagai et al. (2019, 2021). Higher-frequency (initial |kU| > 2.5f) lee waves are confined above the seamount, reflecting from the surface or base of the mixed layer to return to the summit where they may break or impact lee-wave generation (Baker and Mashayek 2021). Since none of the lee waves reach 400-m depth, the elevated turbulence in this layer (Fig. 6) must have another cause such as instabilities in the wake or anisotropic stratified turbulence.
(a) Lee-wave ray paths originating from Hirase summit. Upward-propagating waves reflect off the surface or base of the mixed layer where |kU(zN)| = N(zN) then propagate downward to encounter vertical critical layers as |kU(zf)| → f at depths of 200–300 m (red). Downward-propagating lee waves directly encounter critical layers (blue). (b) Polynomial fits to pycnocline buoyancy frequency (green) and along-stream flow (red) profiles used for the ray tracing.
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-22-0242.1
6. Vertical wavenumber spectra
EM-APEX float vertical wavenumber spectra for normalized shear
Composite vertical wavenumber spectra for (a) normalized shear, (b) vertical strain, and (c) vertical divergence from 200-m half-overlapping EM-APEX profile segments below the mixed layer binned by depth-averaged dissipation rate
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-22-0242.1
Shear and strain spectra are an order of magnitude above the canonical GM spectral level (Garrett and Munk 1975; Cairns and Williams 1976; Gregg and Kunze 1991) (Figs. 11a,b). They have spectral slopes of −1 for wavenumbers m > 0.01 cpm, and their spectral levels show no dependence on turbulence intensity, indicating saturation for which vertical shears |vz| ∼ N (Dewan 1979; Gargett et al. 1981). Spectral roll-offs are at slightly higher wavenumber for strain (mc ≈ 0.02 cpm) than shear (mc ≈ 0.01 cpm). Thus, flows are marginally unstable (Smyth 2020) on wavelengths of
Vertical divergence spectra are one to two orders of magnitude higher than the canonical GM level, with |wz| ∼ N for the strongest turbulence (Fig. 11c). In contrast to the shear and strain spectra, wz spectral levels increase with turbulent intensity and the spectra do not roll off for strongest turbulence. The latter could be because (i) vertical divergence is dominated by near-buoyancy frequency waves (Desaubies 1975) in contrast to shear and strain spectra dominated by near-inertial waves, or (ii) a turbulence contribution. Consistent with (i), Sherman and Pinkel (1991) found that the roll-off wavenumber increased with increasing frequency.
The finescale ratio of horizontal kinetic to available potential energy HKE/APE is near the GM value of 3 at low vertical wavenumbers, decreasing to 1 at higher wavenumbers, regardless of dissipation rate (Fig. 12a). The finescale ratio of horizontal to vertical kinetic energy HKE/VKE is close to the GM value of ∼80 at low wavenumbers and lower turbulent intensities but drops precipitously at high vertical wavenumbers for the strongest turbulence, where flows become almost isotropic (Fig. 12b). The decreases at higher wavenumber could be due to (i) superinertial internal waves (Desaubies 1975) or (ii) turbulence.
Ratios of vertical wavenumber spectra of (a) horizontal kinetic energy HKE to available potential energy APE and (b) HKE to vertical kinetic energy VKE from composite spectra in Fig. 11. Colors denote different dissipation rates
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-22-0242.1
At lower vertical wavenumbers (<0.02 cpm), observed vertical divergence spectra are roughly consistent with GM (11) spectral levels and shapes while model ANISO (13) turbulence has much lower spectral levels (Fig. 13). At higher vertical wavenumbers and lower measured ε < 1.3 × 10−7 W kg−1, observed spectra roll off below GM with levels consistent with ANISO model spectra at the highest vertical wavenumber. At high vertical wavenumbers and the highest measured ε ∼ 7.7 × 10−7 W kg−1, measured spectra are consistent with both GM and the ANISO/ISO turbulence transition which cannot be distinguished. We interpret this figure as signifying that vertical wavenumbers less than a few times 10−2 cpm are linear internal waves while turbulence dominates wavenumbers
Comparison of vertical wavenumber spectra for vertical divergence with Garrett–Munk internal-wave (11) (dashed line), isotropic (12) (solid line at high wavenumbers), and anisotropic turbulence (13) (solid line at low wavenumbers) models. Colors denote different dissipation rates
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-22-0242.1
To summarize this section, observed vertical wavenumber spectra of shear, strain, and vertical divergence suggest 10–100-m vertical scales correspond to the transition between internal waves and turbulence. Vertical divergence spectral levels are compatible with GM levels at ∼10−2 cpm, even though this scale is in the −1 spectral slope regime for shear and strain. The ANISO turbulence model lies below measured vertical divergence at low wavenumbers but becomes comparable at the highest vertical wavenumbers. Finer vertical resolution measurements are needed for more definitive conclusions.
7. Comparison with turbulent parameterizations
In this section, the validity of three turbulence parameterizations is examined for the energetic turbulence observed in Tokara Strait—finescale (Gregg 1989; Polzin et al. 1995), large eddy (Moum 1996; D’Asaro and Lien 2000b), and reduced shear (Kunze et al. 1990).
a. Finescale parameterization
Shear and strain variances are usually computed by integrating vertical wavenumber spectra over the wavenumber band below spectral roll-offs for comparison with GM variances (Gregg and Kunze 1991). Since measured spectra in Tokara Strait have roll-offs on length scales comparable to the water depth, vertical profiles of ε cannot be obtained by this method. As a compromise, first-difference shear and strain are calculated from horizontal velocity and density profiles vertically smoothed using a 50-m cutoff low-pass filter. The 50-m shear and strain variances are averaged over a moving 50-m window and substituted into (14) following Gregg (1989). The 50-m first differences resolve wavenumbers less than 0.1 cpm, which are outside the saturated wave band and in the weakly nonlinear wave regime, so the finescale parameterization should be valid (Gargett 1990; Gregg and Kunze 1991; Polzin et al. 2014b).
Finescale εFS and simultaneously obtained microstructure εmicro have low correlation coefficient R = 0.17 (Fig. 14a), as expected for the finescale parameterization which was never intended to capture individual events (Henyey et al. 1986; Gregg 1989; Polzin et al. 1995; Kunze et al. 2006; Polzin et al. 2014b). However, ensemble-average vertical profiles of finescale 〈εFS〉, as appropriate for this bulk parameterization, and microstructure 〈εmicro〉 are correlated with R = 0.62 (Fig. 15). Finescale estimates 〈εFS〉 are as much as 3 times smaller than microstructure 〈εmicro〉. This may be due to contamination by the saturated vertical wavenumber spectra, which will lead to finescale underestimation of ε because the spectral level is normalized by the GM value in (14) (Gargett 1990; Gregg and Kunze 1991).
Scatterplots of turbulent kinetic energy dissipation rates from the (a) finescale parameterization εFS and microstructure-inferred εmicro and (b) large-eddy parameterization εLE and εmicro. Each estimate is smoothed over a 52-m vertical window to compare all estimates consistently. Colors indicate data number in each bin. Solid diagonal lines mark one-to-one correspondence, and dashed lines mark a factor-of-3 agreement. Correlation coefficients R between two estimates are shown in each panel.
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-22-0242.1
Vertical profiles of ensemble-average dissipation rates 〈ε〉 from the finescale parameterization (FS; blue), large-eddy parameterization (LE; red), reduced-shear parameterization (RS; green), and microstructure measurements (micro; black) for the (a) western and (b) eastern trajectories, respectively. To see downstream conditions more clearly, data north of 30.05°N in the western trajectory and 30°N in the eastern trajectory are excluded from the comparison. Each ensemble-average profile is smoothed over 16-m depth intervals to preserve vertical structure. The 95% confidence intervals of the ensemble-average dissipation rates are not visible because they are smaller than the linewidth. Black vertical bars on the left axes indicate the approximate height of Hirase.
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-22-0242.1
b. Large-eddy parameterization
Oceanic Lagrangian frequency spectra of vertical velocity exhibit a sharp drop at
In analyses by Beaird et al. (2012), mO ∼ 0.2 cpm, m1 = 1/30 cpm, and m2 ∼ 0.5 cpm, i.e., n1 ∼ 0.17 and n2 ∼ 2.5, so c = 0.51 is expected. This value is close to, but above, their least squares fitting c = 0.37. In analyses by Moum (1996), mO ∼ 0.5 cpm, m1 ∼ 1/3.75 cpm, and m2 ∼ 10 cpm, i.e., n1 ∼ 0.53 and n2 ∼ 20. The predicted coefficient c = 1.0 is also close to their least squares fitting c = 0.73. In our data, mO ∼ 1 cpm (Fig. 11), m1 ∼ 1/30 cpm, and m2 ∼ 1/8 cpm, i.e., n1 ∼ 1/3 and n2 ∼ 1.25, so c = 0.87 is expected, which is about a half of our least squares fitting c = 1.9. Our analysis may require c greater than (20) because the energy-containing scale of turbulence (∼the Ozmidov scale) was not always resolved.
Ensemble-mean profiles obtained from the large-eddy parameterization 〈εLE〉 are reasonably consistent with microstructure 〈εmicro〉 with R = 0.67 (Fig. 15). The parameterization tends to overestimate at 100–300-m depths in the western deployment where shear and turbulent dissipation are enhanced, while underestimating below 300 m in the eastern deployment where shear and dissipation are modest (Figs. 8 and 15). This may be explained by our fixed value of c being biased large to account for turbulent vertical velocity variances not resolved by EM-APEX floats for weaker turbulence.
c. Reduced-shear parameterization
Kunze et al. (1990) proposed a parameterization based on reduced shear |vz| − 2N which is applicable for data with sufficient vertical resolution and number of observations to capture the statistics of shear instability. The parameterization assumes that (i) temporally or spatially averaged turbulent dissipation is controlled by intermittent shear instability, (ii) instabilities extract the amount of energy needed to decrease the reduced shear |vz| − 2N below zero, a shear-stable state (Thorpe 1973; Thompson 1980), and (iii) instability growth rate is that of KH billows. It has been validated in the equatorial Pacific (Peters et al. 1995), at midlatitude (Polzin 1996) and in the Columbia River plume (Jurisa et al. 2016). For the present data, application of this parameterization is promising because EM-APEX floats capture shear-unstable events with reduced shear |vz| − 2N > 0 in 24% of the total available profiles (Figs. 8a,b and 16b,d).
Vertical profiles of the components of the reduced-shear parameterization (21) for the (bottom) eastern and (top) western trajectories: (a),(c) fraction of shear-unstable data (red) and ensemble-average thickness of unstable shear layers δh (black) and (b),(d) ensemble-average reduced-shear magnitude
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-22-0242.1
The reduced-shear parameterization can be applied only to shear-unstable layers which are intermittent and not always resolved. This is contrary to the finescale parameterization which is applied to larger scales by assuming the energy cascade to microscale by different processes, weak wave–wave interactions. Because only a limited number of unstable events are captured, we focus on the ensemble-average vertical profiles to compare the reduced-shear estimates with those from microstructure and the other parameterizations.
Parameterization (21) is applied to EM-APEX horizontal velocity and density profiles to obtain 〈ε〉RS as a function of depth for the western and eastern trajectories for comparison with average microstructure profiles 〈ε〉micro. The 〈ε〉RS is only calculated below the mixed layer (z > 100 m) and in the 4-m depth intervals where the total number of available profiles exceeds 50.
Reduced-shear and microstructure estimates of 〈ε〉 are consistent (Fig. 15), suggesting that EM-APEX floats satisfactorily capture unstable shear layers in Tokara Strait. The correlation R = 0.80 is higher than for large-eddy (R = 0.67) and finescale (R = 0.62) parameterizations. In the western trajectory, unstable shears were most frequently observed at ∼150- and ∼250-m depths with 5–10-m thicknesses ∼1.5–2 times thinner than in the deeper layer (Fig. 16a). In the eastern trajectory, unstable shears were frequently observed at ∼200- and ∼400-m depths, where they have similar 10–15-m thicknesses (Fig. 16c). In both trajectories, KH growth rates are ∼2 times higher and available unstable energy more than 4 times greater in the shallower layer than in the deeper layer, which results in more elevated turbulent dissipation in the shallower layer (Figs. 16b,d).
8. Conclusions
EM-APEX profiling float and ADCP mooring observations were used to examine energetic fine- and microstructure downstream of ∼10-km-wide Hirase Seamount in the path of the Kuroshio in Tokara Strait (Fig. 1). Horizontal and vertical scales of Hirase Seamount, background velocity (Fig. 2), and stratification (Fig. 3) suggest that upstream flows will split around the seamount and vertically sheared vortices will form in the downstream wake of the seamount flanks (e.g., Baines 1995; Srinivasan et al. 2021), consistent with EM-APEX float profile sections (Fig. 4). Conventional upgoing lee-wave generation will be confined to the summit crown. Mooring frequency spectra of vertical shear have broadband spectral enhancement from subinertial to supertidal at 100–300-m depth downstream (Fig. 5).
The EM-APEX floats capture energetic turbulence above the seamount flanks and extending downstream (i) for at least 20 km in a layer spanning 150–250-m depth in the pycnocline where TKE dissipation rates
Previous measurements indicate that layers of elevated turbulence extend at least 100 km downstream of Hirase (Nagai et al. 2017, 2019, 2021; Hasegawa et al. 2021). That turbulence extends farther downstream than expected for turbulent decay times
Vertical wavenumber spectra for vertical shear, strain, and vertical divergence (Fig. 11) are orders of magnitude higher than canonical GM levels (Garrett and Munk 1975; Cairns and Williams 1976; Gregg and Kunze 1991; Thurnherr et al. 2015). Shear and strain spectra are saturated with spectral levels invariant with ε, |vz| ∼ N, and spectral slopes of ∼−1 for vertical wavenumbers as low as 0.01 cpm, suggesting that flows are nonlinear on scales comparable to water depths (Figs. 11a,b). In contrast, vertical divergence spectral levels increase with increasing ε (Fig. 11c). Spectral amplitudes of vertical kinetic energy are comparable to those of horizontal kinetic energy at 10-m vertical scales for the strongest dissipation rates, signifying isotropy and consistent with the measurements resolving the Ozmidov length (Fig. 12). Observed vertical divergence wz is consistent with the GM model (Thurnherr et al. 2015) at 100-m vertical scales, while agreeing with the transition between anisotropic turbulence (Kunze 2019) and isotropic turbulence (Kolmogorov 1941; Batchelor 1953; Ozmidov 1965; Tennekes and Lumley 1972; Thorpe 2005) at 10-m vertical scales (Fig. 13).
Measured ε correlates well with finescale (Gregg 1989; Polzin et al. 1995; Gregg et al. 2003), large-eddy (Moum 1996; D’Asaro and Lien 2000b), and reduced-shear (Kunze et al. 1990) parameterizations ranging from the roll-off length scales of weakly nonlinear internal waves (∼100 m in this case) to the Ozmidov length scales of isotropic turbulence (∼10 m for the strongest turbulence) (Figs. 14–16). The large-eddy scaling appears to work even though the Ozmidov scales are not fully resolved. The lee of Hirase is close to the wave–turbulence transition (D’Asaro and Lien 2000b) with the saturated spectrum (anisotropic stratified turbulence) extending to vertical length scales as large as 100 m, comparable to the water depth in Tokara Strait.
Extensive EM-APEX float measurements in this study demonstrate relationships between microstructure and vertical fine structure. Flow–topography interaction theory suggests the vertical fine structure contains lee waves and submesoscale vortical motions with similar temporal and spatial scales. The two could not be distinguished in the present 2D measurements using the float’s line array because PV could not be estimated. Future 3D measurements and numerical simulations may allow identification of turbulence generation mechanisms on the seamount flanks and downstream more rigorously. Time-evolving tidal currents may make the Kuroshio–seamount interactions more complex (e.g., MacKinnon et al. 2019; Inoue et al. 2024), which could be addressed with 3D numerical simulations and long-term mooring observations.
Acknowledgments.
The authors are grateful to the crew and officers of the TRV Kagoshima-maru, as well as all November 2019 cruise participants for their assistance. Special thanks go to Avery Snyder, Takeshi Matsuno, Toru Kobari, and Akie Sakai. Sebastian Essink helped with FP07 data analysis. We also appreciate an anonymous reviewer for insightful comments. This work was supported by the National Science Foundation Grants OCE-1829082 and OCE-1829190 and Japan Society for the Promotion of Science KAKENHI Grants JP15H05818 and JP15H05821.
Data availability statement.
The EM-APEX float and ADCP mooring data are available at https://hdl.handle.net/1773/49468. ETOPO1 bathymetry data are distributed by NOAA National Geophysical Data Center (https://www.ncei.noaa.gov/access/metadata/landing-page/bin/iso?id=gov.noaa.ngdc.mgg.dem:316). Global Ocean Physics Reanalysis data are distributed by CMEMS (https://resources.marine.copernicus.eu/product-detail/GLOBAL_MULTIYEAR_PHY_001_030).
APPENDIX A
Background Density Profiles from EM-APEX Floats
Stable background density profiles for the western and eastern float trajectories are constructed separately, using all EM-APEX float profiles along these two trajectories. For each deployment, a total of Ntot density points were measured from all floats with vertical resolution Δz = 2 m to maximum depth Hmax = MΔz. For each depth interval (i − 1)Δz < z < iΔz (i = 2, …, M), a finer vertical interval is defined δzi = Δz/ni, where ni is the total number of data points in this depth interval such that
APPENDIX B
Quality Control of Microstructure Data
Comparison between measured (solid lines) and Kraichnan universal temperature gradient spectrum (dotted lines), both averaged for data with dissipation rates ε and χ shown above the upper axes. Dashed vertical lines indicate minimum (
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-22-0242.1
APPENDIX C
Ray Tracing of Lee Waves in the Kuroshio Shear
In (C5), vertical increment is set as dz = 1 m if the resulting horizontal increment dx < 10 m, otherwise horizontal increment dx = 10 m is used. Rays are terminated (i) if they fall 20 m below the bottom (gray shading in Fig. 10a), (ii) if they leave the domain (x > 20 km), or (iii) if the vertical wavenumber magnitude |m| exceeds 0.1 cpm.
REFERENCES
Althaus, A. M., E. Kunze, and T. B. Sanford, 2003: Internal tide radiation from Mendocino escarpment. J. Phys. Oceanogr., 33, 1510–1527, https://doi.org/10.1175/1520-0485(2003)033<1510:ITRFME>2.0.CO;2.
Baines, P. G., 1995: Topographic Effects in Stratified Flows. Cambridge University Press, 482 pp.
Baker, L., and A. Mashayek, 2021: Surface reflection of bottom generated oceanic lee waves. J. Fluid Mech., 924, A17, https://doi.org/10.1017/jfm.2021.627.
Balmforth, N. J., G. R. Ierley, and W. R. Young, 2002: Tidal conversion by subcritical topography. J. Phys. Oceanogr., 32, 2900–2914, https://doi.org/10.1175/1520-0485(2002)032%3C2900:TCBST%3E2.0.CO;2.
Batchelor, G. K., 1953: The Theory of Homogeneous Turbulence. Cambridge University Press, 197 pp.
Batchelor, G. K., 1959: Small-scale variation of convected quantities like temperature in turbulent fluid Part 1. General discussion and the case of small conductivity. J. Fluid Mech., 5, 113–133, https://doi.org/10.1017/S002211205900009X.
Beaird, N., I. Fer, P. Rhines, and C. Eriksen, 2012: Dissipation of turbulent kinetic energy inferred from Seagliders: An application to the eastern Nordic Seas overflows. J. Phys. Oceanogr., 42, 2268–2282, https://doi.org/10.1175/JPO-D-12-094.1.
Bell, T. H., 1975a: Lee waves in stratified flows with simple harmonic time dependence. J. Fluid Mech., 67, 705–722, https://doi.org/10.1017/S0022112075000560.
Bell, T. H., Jr., 1975b: Topographically generated internal waves in the open ocean. J. Geophys. Res., 80, 320–327, https://doi.org/10.1029/JC080i003p00320.
Bogucki, D., J. A. Domaradzki, and P. K. Yeung, 1997: Direct numerical simulations of passive scalars with Pr>1 advected by turbulent flow. J. Fluid Mech., 343, 111–130, https://doi.org/10.1017/S0022112097005727.
Bogucki, D. J., H. Luo, and J. A. Domaradzki, 2012: Experimental evidence of the Kraichnan scalar spectrum at high Reynolds numbers. J. Phys. Oceanogr., 42, 1717–1728, https://doi.org/10.1175/JPO-D-11-0214.1.
Burchard, H., O. Petersen, and T. P. Rippeth, 1998: Comparing the performance of the Mellor-Yamada and the κ-ϵ two-equation turbulence models. J. Geophys. Res., 103, 10 543–10 554, https://doi.org/10.1029/98JC00261.
Cairns, J. L., and G. O. Williams, 1976: Internal wave observations from a midwater float, 2. J. Geophys. Res., 81, 1943–1950, https://doi.org/10.1029/JC081i012p01943.
Caldeira, R. M. A., A. Stegner, X. Couvelard, I. B. Araújo, P. Testor, and A. Lorenzo, 2014: Evolution of an oceanic anticyclone in the lee of Madeira Island: In situ and remote sensing survey. J. Geophys. Res. Oceans, 119, 1195–1216, https://doi.org/10.1002/2013JC009493.
Carter, G. S., and M. C. Gregg, 2002: Intense, variable mixing near the head of Monterey Submarine Canyon. J. Phys. Oceanogr., 32, 3145–3165, https://doi.org/10.1175/1520-0485(2002)032<3145:IVMNTH>2.0.CO;2.
Chang, M.-H., T. Y. Tang, C.-R. Ho, and S.-Y. Chao, 2013: Kuroshio-induced wake in the lee of Green Island off Taiwan. J. Geophys. Res. Oceans, 118, 1508–1519, https://doi.org/10.1002/jgrc.20151.
Chang, M.-H., S.-Y. Jheng, and R.-C. Lien, 2016: Trains of large Kelvin-Helmholtz billows observed in the Kuroshio above a seamount. Geophys. Res. Lett., 43, 8654–8661, https://doi.org/10.1002/2016GL069462.
Chang, M.-H., S. Jan, C.-L. Liu, Y.-H. Cheng, and V. Mensah, 2019: Observations of island wakes at high Rossby numbers: Evolution of submesoscale vortices and free shear layers. J. Phys. Oceanogr., 49, 2997–3016, https://doi.org/10.1175/JPO-D-19-0035.1.
Codiga, D. L., 2011: Unified tidal analysis and prediction using the UTide Matlab functions. Graduate School of Oceanography–University of Rhode Island Tech. Rep. 2011-01, 59 pp.
Cusack, J. M., A. C. Naveira Garabato, D. A. Smeed, and J. B. Girton, 2017: Observation of a large lee wave in the Drake Passage. J. Phys. Oceanogr., 47, 793–810, https://doi.org/10.1175/JPO-D-16-0153.1.
D’Asaro, E. A., 1988: Generation of submesoscale vortices: A new mechanism. J. Geophys. Res., 93, 6685–6693, https://doi.org/10.1029/JC093iC06p06685.
D’Asaro, E. A., and R.-C. Lien, 2000a: Lagrangian measurements of waves and turbulence in stratified flows. J. Phys. Oceanogr., 30, 641–655, https://doi.org/10.1175/1520-0485(2000)030<0641:LMOWAT>2.0.CO;2.
D’Asaro, E. A., and R.-C. Lien, 2000b: The wave–turbulence transition for stratified flows. J. Phys. Oceanogr., 30, 1669–1678, https://doi.org/10.1175/1520-0485(2000)030%3C1669:TWTTFS%3E2.0.CO;2.
Desaubies, Y. J. F., 1975: A linear theory of internal wave spectra and coherences near the Väisälä frequency. J. Geophys. Res., 80, 895–899, https://doi.org/10.1029/JC080i006p00895.
Dewan, E. M., 1979: Stratospheric wave spectra resembling turbulence. Science, 204, 832–835, https://doi.org/10.1126/science.204.4395.832.
Eriksen, C. C., 1982: Observations of internal wave reflection off sloping bottoms. J. Geophys. Res., 87, 525–538, https://doi.org/10.1029/JC087iC01p00525.
Eriksen, C. C., 1998: Internal wave reflection and mixing at Fieberling Guyot. J. Geophys. Res., 103, 2977–2994, https://doi.org/10.1029/97JC03205.
Essink, S., E. Kunze, R.-C. Lien, R. Inoue, and S.-I. Ito, 2022: Near-inertial wave interactions and turbulence production in a Kuroshio anticyclonic eddy. J. Phys. Oceanogr., 52, 2687–2704, https://doi.org/10.1175/JPO-D-21-0278.1.
Evans, D. G., and Coauthors, 2018: Annual cycle of turbulent dissipation estimated from Seagliders. Geophys. Res. Lett., 45, 10 560–10 569, https://doi.org/10.1029/2018GL079966.
Farmer, D., and L. Armi, 1999: Stratified flow over topography: The role of small-scale entrainment and mixing in flow establishment. Proc. Roy. Soc. London, 455A, 3221–3258, https://doi.org/10.1098/rspa.1999.0448.
Frajka-Williams, E., C. C. Eriksen, P. B. Rhines, and R. R. Harcourt, 2011: Determining vertical water velocities from Seaglider. J. Atmos. Oceanic Technol., 28, 1641–1656, https://doi.org/10.1175/2011JTECHO830.1.
Gargett, A. E., 1990: Do we really know how to scale the turbulent kinetic energy dissipation rate ϵ due to breaking of oceanic internal waves? J. Geophys. Res., 95, 15 971–15 974, https://doi.org/10.1029/JC095iC09p15971.
Gargett, A. E., P. J. Hendricks, T. B. Sanford, T. R. Osborn, and A. J. Williams, 1981: A composite spectrum of vertical shear in the upper ocean. J. Phys. Oceanogr., 11, 1258–1271, https://doi.org/10.1175/1520-0485(1981)011<1258:ACSOVS>2.0.CO;2.
Garrett, C., and W. Munk, 1975: Space-time scales of internal waves: A progress report. J. Geophys. Res., 80, 291–297, https://doi.org/10.1029/JC080i003p00291.
Garrett, C., and W. Munk, 1979: Internal waves in the ocean. Annu. Rev. Fluid Mech., 11, 339–369, https://doi.org/10.1146/annurev.fl.11.010179.002011.
Garrett, C., and E. Kunze, 2007: Internal tide generation in the deep ocean. Annu. Rev. Fluid Mech., 39, 57–87, https://doi.org/10.1146/annurev.fluid.39.050905.110227.
Goto, Y., I. Yasuda, and M. Nagasawa, 2016: Turbulence estimation using fast-response thermistors attached to a free-fall vertical microstructure profiler. J. Atmos. Oceanic Technol., 33, 2065–2078, https://doi.org/10.1175/JTECH-D-15-0220.1.
Gregg, M. C., 1989: Scaling turbulent dissipation in the thermocline. J. Geophys. Res., 94, 9686–9698, https://doi.org/10.1029/JC094iC07p09686.
Gregg, M. C., and E. Kunze, 1991: Shear and strain in Santa Monica Basin. J. Geophys. Res., 96, 16 709–16 719, https://doi.org/10.1029/91JC01385.
Gregg, M. C., T. B. Sanford, and D. P. Winkel, 2003: Reduced mixing from the breaking of internal waves in equatorial waters. Nature, 422, 513–515, https://doi.org/10.1038/nature01507.
Gregg, M. C., E. A. D’Asaro, J. J. Riley, and E. Kunze, 2018: Mixing efficiency in the ocean. Annu. Rev. Mar. Sci., 10, 443–473, https://doi.org/10.1146/annurev-marine-121916-063643.
Guo, X., X.-H. Zhu, Q.-S. Wu, and D. Huang, 2012: The Kuroshio nutrient stream and its temporal variation in the East China Sea. J. Geophys. Res., 117, C01026, https://doi.org/10.1029/2011JC007292.
Guo, X. Y., X.-H. Zhu, Y. Long, and D. J. Huang, 2013: Spatial variations in the Kuroshio nutrient transport from the East China Sea to South of Japan. Biogeosciences, 10, 6403–6417, https://doi.org/10.5194/bg-10-6403-2013.
Hasegawa, D., and Coauthors, 2021: How a small reef in the Kuroshio cultivates the ocean. Geophys. Res. Lett., 48, e2020GL092063, https://doi.org/10.1029/2020GL092063.
Hazel, P., 1972: Numerical studies of the stability of inviscid stratified shear flows. J. Fluid Mech., 51, 39–61, https://doi.org/10.1017/S0022112072001065.
Henyey, F. S., J. Wright, and S. M. Flatté, 1986: Energy and action flow through the internal wave field: An eikonal approach. J. Geophys. Res., 91, 8487–8495, https://doi.org/10.1029/JC091iC07p08487.
Hibiya, T., N. Furuichi, and R. Robertson, 2012: Assessment of fine-scale parameterizations of turbulent dissipation rates near mixing hotspots in the deep ocean. Geophys. Res. Lett., 39, L24601, https://doi.org/10.1029/2012GL054068.
Inoue, R., E. Tsutsumi, and H. Nakamura, 2024: Numerical simulation of the Kuroshio flowing over the Hirase Seamount in the Tokara Strait in autumn: Tidal vortex shedding in a baroclinic jet. J. Phys. Oceanogr., 54, 153–172, https://doi.org/10.1175/JPO-D-23-0050.1.
Johnston, T. M. S., and Coauthors, 2019: Energy and momentum lost to wake eddies and lee waves generated by the North Equatorial Current and tidal flows at Peleliu, Palau. Oceanography, 32, 110–125, https://doi.org/10.5670/oceanog.2019.417.
Jurisa, J. T., J. D. Nash, J. N. Moum, and L. F. Kilcher, 2016: Controls on turbulent mixing in a strongly stratified and sheared tidal river plume. J. Phys. Oceanogr., 46, 2373–2388, https://doi.org/10.1175/JPO-D-15-0156.1.
Klymak, J. M., and M. C. Gregg, 2004: Tidally generated turbulence over the Knight Inlet sill. J. Phys. Oceanogr., 34, 1135–1151, https://doi.org/10.1175/1520-0485(2004)034%3C1135:TGTOTK%3E2.0.CO;2.
Klymak, J. M., R. Pinkel, and L. Rainville, 2008: Direct breaking of the internal tide near topography: Kaena Ridge, Hawaii. J. Phys. Oceanogr., 38, 380–399, https://doi.org/10.1175/2007JPO3728.1.
Kolmogorov, A., 1941: The local structure of turbulence in incompressible viscous fluids for very large Reynolds number. Dokl. Akad. Nauk SSSR, 30, 199–303.
Kraichnan, R. H., 1968: Small-scale structure of a scalar field convected by turbulence. Phys. Fluids, 11, 945–953, https://doi.org/10.1063/1.1692063.
Kunze, E., 1985: Near-inertial wave propagation in geostrophic shear. J. Phys. Oceanogr., 15, 544–565, https://doi.org/10.1175/1520-0485(1985)015%3C0544:NIWPIG%3E2.0.CO;2.
Kunze, E., 2017: Internal-wave-driven mixing: Global geography and budgets. J. Phys. Oceanogr., 47, 1325–1345, https://doi.org/10.1175/JPO-D-16-0141.1.
Kunze, E., 2019: A unified model spectrum for anisotropic stratified and isotropic turbulence in the ocean and atmosphere. J. Phys. Oceanogr., 49, 385–407, https://doi.org/10.1175/JPO-D-18-0092.1.
Kunze, E., and J. M. Toole, 1997: Tidally driven vorticity, diurnal shear, and turbulence atop Fieberling Seamount. J. Phys. Oceanogr., 27, 2663–2693, https://doi.org/10.1175/1520-0485(1997)027<2663:TDVDSA>2.0.CO;2.
Kunze, E., and R.-C. Lien, 2019: Energy sinks for lee waves in shear flow. J. Phys. Oceanogr., 49, 2851–2865, https://doi.org/10.1175/JPO-D-19-0052.1.
Kunze, E., A. J. Williams III, and M. G. Briscoe, 1990: Observations of shear and vertical stability from a neutrally buoyant float. J. Geophys. Res., 95, 18 127–18 142, https://doi.org/10.1029/JC095iC10p18127.
Kunze, E., R. W. Schmitt, and J. M. Toole, 1995: The energy balance in a warm-core ring’s near-inertial critical layer. J. Phys. Oceanogr., 25, 942–957, https://doi.org/10.1175/1520-0485(1995)025<0942:TEBIAW>2.0.CO;2.
Kunze, E., L. K. Rosenfeld, G. S. Carter, and M. C. Gregg, 2002: Internal waves in Monterey Submarine Canyon. J. Phys. Oceanogr., 32, 1890–1913, https://doi.org/10.1175/1520-0485(2002)032<1890:IWIMSC>2.0.CO;2.
Kunze, E., E. Firing, J. M. Hummon, T. K. Chereskin, and A. M. Thurnherr, 2006: Global abyssal mixing inferred from lowered ADCP shear and CTD strain profiles. J. Phys. Oceanogr., 36, 1553–1576, https://doi.org/10.1175/JPO2926.1.
Kunze, E., C. MacKay, E. E. McPhee-Shaw, K. Morrice, J. B. Girton, and S. R. Terker, 2012: Turbulent mixing and exchange with interior waters on sloping boundaries. J. Phys. Oceanogr., 42, 910–927, https://doi.org/10.1175/JPO-D-11-075.1.
Legg, S., 2021: Mixing by oceanic lee waves. Annu. Rev. Fluid Mech., 53, 173–201, https://doi.org/10.1146/annurev-fluid-051220-043904.
Lien, R.-C., and T. B. Sanford, 2019: Small-scale potential vorticity in the upper-ocean thermocline. J. Phys. Oceanogr., 49, 1845–1872, https://doi.org/10.1175/JPO-D-18-0052.1.
Lien, R.-C., M. J. McPhaden, and M. C. Gregg, 1996: High-frequency internal waves at 0°, 140°W and their possible relationship to deep-cycle turbulence. J. Phys. Oceanogr., 26, 581–600, https://doi.org/10.1175/1520-0485(1996)026<0581:HFIWAA>2.0.CO;2.
Lien, R.-C., T. B. Sanford, J. A. Carlson, and J. H. Dunlap, 2016: Autonomous microstructure EM-APEX floats. Methods Oceanogr., 17, 282–295, https://doi.org/10.1016/j.mio.2016.09.003.
Lighthill, M. J., 1978: Waves in Fluids. Cambridge University Press, 504 pp.
MacCready, P., and P. B. Rhines, 1993: Slippery bottom boundary layers on a slope. J. Phys. Oceanogr., 23, 5–22, https://doi.org/10.1175/1520-0485(1993)023%3C0005:SBBLOA%3E2.0.CO;2.
MacKinnon, J. A., and M. C. Gregg, 2005: Spring mixing: Turbulence and internal waves during restratification on the New England Shelf. J. Phys. Oceanogr., 35, 2425–2443, https://doi.org/10.1175/JPO2821.1.
MacKinnon, J. A., M. H. Alford, G. Voet, K. L. Zeiden, T. M. Shaun Johnston, M. Siegelman, S. Merrifield, and M. Merrifield, 2019: Eddy wake generation from broadband currents near Palau. J. Geophys. Res. Oceans, 124, 4891–4903, https://doi.org/10.1029/2019JC014945.
McComas, C. H., and P. Müller, 1981: The dynamic balance of internal waves. J. Phys. Oceanogr., 11, 970–986, https://doi.org/10.1175/1520-0485(1981)011%3C0970:TDBOIW%3E2.0.CO;2.
McDougall, T. J., and P. Barker, 2011: Getting started with TEOS-10 and the Gibbs Seawater (GSW) Oceanographic Toolbox. SCOR/IAPSO WG127, 28 pp., https://www.teos-10.org/pubs/Getting_Started.pdf.
Mellor, G. L., and T. Yamada, 1982: Development of a turbulence closure model for geophysical fluid problems. Rev. Geophys., 20, 851–875, https://doi.org/10.1029/RG020i004p00851.
Mohri, K., T. Hibiya, and N. Iwamae, 2010: Revisiting internal wave generation by tide-topography interaction. J. Geophys. Res., 115, C11001, https://doi.org/10.1029/2009JC005908.
Molemaker, M. J., J. C. McWilliams, and W. K. Dewar, 2015: Submesoscale instability and generation of mesoscale anticyclones near a separation of the California Undercurrent. J. Phys. Oceanogr., 45, 613–629, https://doi.org/10.1175/JPO-D-13-0225.1.
Monismith, S. G., J. R. Koseff, and B. L. White, 2018: Mixing efficiency in the presence of stratification: When is it constant? Geophys. Res. Lett., 45, 5627–5634, https://doi.org/10.1029/2018GL077229.
Moum, J. N., 1996: Energy-containing scales of turbulence in the ocean thermocline. J. Geophys. Res., 101, 14 095–14 109, https://doi.org/10.1029/96JC00507.
Moum, J. N., and J. D. Nash, 2000: Topographically induced drag and mixing at a small bank on the continental shelf. J. Phys. Oceanogr., 30, 2049–2054, https://doi.org/10.1175/1520-0485(2000)030<2049:TIDAMA>2.0.CO;2.
Müller, P., and N. Xu, 1992: Scattering of oceanic internal gravity waves off random bottom topography. J. Phys. Oceanogr., 22, 474–488, https://doi.org/10.1175/1520-0485(1992)022<0474:SOOIGW>2.0.CO;2.
Nagai, T., D. Hasegawa, T. Tanaka, H. Nakamura, E. Tsutsumi, R. Inoue, and T. Yamashiro, 2017: First evidence of coherent bands of strong turbulent layers associated with high-wavenumber internal-wave shear in the upstream Kuroshio. Sci. Rep., 7, 14555, https://doi.org/10.1038/s41598-017-15167-1.
Nagai, T., and Coauthors, 2019: How the Kuroshio Current delivers nutrients to sunlit layers on the continental shelves with aid of near-inertial waves and turbulence. Geophys. Res. Lett., 46, 6726–6735, https://doi.org/10.1029/2019GL082680.
Nagai, T., and Coauthors, 2021: The Kuroshio flowing over seamounts and associated submesoscale flows drive 100-km-wide 100-1000-fold enhancement of turbulence. Commun. Earth Environ., 2, 170, https://doi.org/10.1038/s43247-021-00230-7.
Nash, J. D., E. Kunze, J. M. Toole, and R. W. Schmitt, 2004: Internal tide reflection and turbulent mixing on the continental slope. J. Phys. Oceanogr., 34, 1117–1134, https://doi.org/10.1175/1520-0485(2004)034<1117:ITRATM>2.0.CO;2.
Nash, J. D., E. Kunze, C. M. Lee, and T. B. Sanford, 2006: Structure of the baroclinic tide generated at Kaena Ridge, Hawaii. J. Phys. Oceanogr., 36, 1123–1135, https://doi.org/10.1175/JPO2883.1.
Nasmyth, P. W., 1970: Oceanic turbulence. Ph.D. thesis, University of British Columbia, 69 pp.
Naveira Garabato, A. C., K. L. Polzin, B. A. King, K. J. Heywood, and M. Visbeck, 2004: Widespread intense turbulent mixing in the Southern Ocean. Science, 303, 210–213, https://doi.org/10.1126/science.1090929.
Nikurashin, M., and R. Ferrari, 2010: Radiation and dissipation of internal waves generated by geostrophic motions impinging on small-scale topography: Theory. J. Phys. Oceanogr., 40, 1055–1074, https://doi.org/10.1175/2009JPO4199.1.
Nikurashin, M., R. Ferrari, N. Grisouard, and K. Polzin, 2014: The impact of finite-amplitude bottom topography on internal wave generation in the Southern Ocean. J. Phys. Oceanogr., 44, 2938–2950, https://doi.org/10.1175/JPO-D-13-0201.1.
Oakey, N. S., 1982: Determination of the rate of dissipation of turbulent energy from simultaneous temperature and velocity shear microstructure measurements. J. Phys. Oceanogr., 12, 256–271, https://doi.org/10.1175/1520-0485(1982)012<0256:DOTROD>2.0.CO;2.
Osborn, T. R., 1980: Estimates of the local rate of vertical diffusion from dissipation measurements. J. Phys. Oceanogr., 10, 83–89, https://doi.org/10.1175/1520-0485(1980)010%3C0083:EOTLRO%3E2.0.CO;2.
Osborn, T. R., and C. S. Cox, 1972: Oceanic fine structure. Geophys. Fluid Dyn., 3, 321–345, https://doi.org/10.1080/03091927208236085.
Ozmidov, R. V., 1965: On the turbulent exchange in a stably stratified ocean. Izv. Acad. Sci. USSR, Atmos. Oceanic Phys., 1, 853–860.
Perfect, B., N. Kumar, and J. J. Riley, 2018: Vortex structures in the wake of an idealized seamount in rotating, stratified flow. Geophys. Res. Lett., 45, 9098–9105, https://doi.org/10.1029/2018GL078703.
Perfect, B., N. Kumar, and J. J. Riley, 2020a: Energetics of seamount wakes. Part I: Energy exchange. J. Phys. Oceanogr., 50, 1365–1382, https://doi.org/10.1175/JPO-D-19-0105.1.
Perfect, B., N. Kumar, and J. J. Riley, 2020b: Energetics of seamount wakes. Part II: Wave fluxes. J. Phys. Oceanogr., 50, 1383–1398, https://doi.org/10.1175/JPO-D-19-0104.1.
Peters, H., M. C. Gregg, and T. B. Sanford, 1995: On the parameterization of equatorial turbulence: Effect of fine-scale variations below the range of the diurnal cycle. J. Geophys. Res., 100, 18 333–18 348, https://doi.org/10.1029/95JC01513.
Peterson, A. K., and I. Fer, 2014: Dissipation measurements using temperature microstructure from an underwater glider. Methods Oceanogr., 10, 44–69, https://doi.org/10.1016/j.mio.2014.05.002.
Polzin, K., 1996: Statistics of the Richardson number: Mixing models and finestructure. J. Phys. Oceanogr., 26, 1409–1425, https://doi.org/10.1175/1520-0485(1996)026%3C1409:SOTRNM%3E2.0.CO;2.
Polzin, K. L., J. M. Toole, and R. W. Schmitt, 1995: Finescale parameterizations of turbulent dissipation. J. Phys. Oceanogr., 25, 306–328, https://doi.org/10.1175/1520-0485(1995)025<0306:FPOTD>2.0.CO;2.
Polzin, K. L., J. M. Toole, J. R. Ledwell, and R. W. Schmitt, 1997: Spatial variability of turbulent mixing in the abyssal ocean. Science, 276, 93–96, https://doi.org/10.1126/science.276.5309.93.
Polzin, K. L., E. Kunze, J. Hummon, and E. Firing, 2002: The finescale response of lowered ADCP velocity profiles. J. Atmos. Oceanic Technol., 19, 205–224, https://doi.org/10.1175/1520-0426(2002)019<0205:TFROLA>2.0.CO;2.
Polzin, K. L., A. C. Naveira Garabato, E. P. Abrahamsen, L. Jullion, and M. P. Meredith, 2014a: Boundary mixing in Orkney Passage outflow. J. Geophys. Res. Oceans, 119, 8627–8645, https://doi.org/10.1002/2014JC010099.
Polzin, K. L., A. C. Naveira Garabato, T. N. Huussen, B. M. Sloyan, and S. Waterman, 2014b: Finescale parameterizations of turbulent dissipation. J. Geophys. Res. Oceans, 119, 1383–1419, https://doi.org/10.1002/2013JC008979.
Puthan, P., S. Sarkar, and G. Pawlak, 2021: Tidal synchronization of lee vortices in geophysical wakes. Geophys. Res. Lett., 48, e2020GL090905, https://doi.org/10.1029/2020GL090905.
Roget, E., I. Lozovatsky, X. Sanchez, and M. Figueroa, 2006: Microstructure measurements in natural waters: Methodology and applications. Prog. Oceanogr., 70, 126–148, https://doi.org/10.1016/j.pocean.2006.07.003.
Sanford, T. B., R. G. Drever, and J. H. Dunlap, 1978: A velocity profiler based on the principles of geomagnetic induction. Deep-Sea Res., 25, 183–210, https://doi.org/10.1016/0146-6291(78)90006-1.
Sanford, T., J. Dunlap, J. Carlson, D. Webb, and J. Girton, 2005: Autonomous velocity and density profiler: EM-APEX. Proc. IEEE/OES Eighth Working Conf. on Current Measurement Technology, Southampton, United Kingdom, IEEE, 152–156, https://doi.org/10.1109/CCM.2005.1506361.
Scott, R. B., and Y. Xu, 2009: An update on the wind power input to the surface geostrophic flow of the world ocean. Deep-Sea Res. I, 56, 295–304, https://doi.org/10.1016/j.dsr.2008.09.010.
Sherman, J. T., and R. Pinkel, 1991: Estimates of the vertical wavenumber–frequency spectra of vertical shear and strain. J. Phys. Oceanogr., 21, 292–303, https://doi.org/10.1175/1520-0485(1991)021<0292:EOTVWS>2.0.CO;2.
Smyth, W. D., 2020: Marginal instability and the efficiency of ocean mixing. J. Phys. Oceanogr., 50, 2141–2150, https://doi.org/10.1175/JPO-D-20-0083.1.
Srinivasan, K., J. C. McWilliams, M. J. Molemaker, and R. O. Barkan, 2019: Submesoscale vortical wakes in the lee of topography. J. Phys. Oceanogr., 49, 1949–1971, https://doi.org/10.1175/JPO-D-18-0042.1.
Srinivasan, K., J. C. McWilliams, and A. Jagannathan, 2021: High vertical shear and dissipation in equatorial topographic wakes. J. Phys. Oceanogr., 51, 1985–2001, https://doi.org/10.1175/JPO-D-20-0119.1.
St. Laurent, L., S. Stringer, C. Garrett, and D. Perrault-Joncas, 2003: The generation of internal tides at abrupt topography. Deep-Sea Res. I, 50, 987–1003, https://doi.org/10.1016/S0967-0637(03)00096-7.
Takahashi, T., and Coauthors, 2009: Climatological mean and decadal change in surface ocean pCO2, and net sea–air CO2 flux over the global oceans. Deep-Sea Res. II, 56, 554–577, https://doi.org/10.1016/j.dsr2.2008.12.009.
Taylor, G. I., 1935: Statistical theory of turbulence. Proc. Roy. Soc. London, 151A, 421–444, https://doi.org/10.1098/rspa.1935.0158.
Tennekes, H., and J. L. Lumley, 1972: A First Course in Turbulence. MIT Press, 320 pp.
Thomas, L. N., J. R. Taylor, R. Ferrari, and T. M. Joyce, 2013: Symmetric instability in the Gulf Stream. Deep-Sea Res. II, 91, 96–110, https://doi.org/10.1016/j.dsr2.2013.02.025.
Thompson, R. O. R. Y., 1980: Efficiency of conversion of kinetic energy to potential energy by a breaking internal gravity wave. J. Geophys. Res., 85, 6631–6635, https://doi.org/10.1029/JC085iC11p06631.
Thomson, R. E., and W. J. Emery, 2014: Time series analysis methods. Data Analysis Methods in Physical Oceanography, 3rd ed. R. E. Thomson and W. J. Emery, Eds., Elsevier, 425–591.
Thorpe, S. A., 1973: Turbulence in stably stratified fluids: A review of laboratory experiments. Bound.-Layer Meteor., 5, 95–119, https://doi.org/10.1007/BF02188314.
Thorpe, S. A., 1992: The generation of internal waves by flow over the rough topography of a continental slope. Proc. Roy. Soc. London, 439A, 115–130, https://doi.org/10.1098/rspa.1992.0137.
Thorpe, S. A., 2005: The Turbulent Ocean. Cambridge University Press, 484 pp.
Thurnherr, A. M., E. Kunze, J. M. Toole, L. St. Laurent, K. J. Richards, and A. Ruiz-Angulo, 2015: Vertical kinetic energy and turbulent dissipation in the ocean. Geophys. Res. Lett., 42, 7639–7647, https://doi.org/10.1002/2015GL065043.
Toole, J. M., R. W. Schmitt, K. L. Polzin, and E. Kunze, 1997: Near-boundary mixing above the flanks of a midlatitude seamount. J. Geophys. Res., 102, 947–959, https://doi.org/10.1029/96JC03160.
Trowbridge, J. H., and S. J. Lentz, 1991: Asymmetric behavior of an oceanic boundary layer above a sloping bottom. J. Phys. Oceanogr., 21, 1171–1185, https://doi.org/10.1175/1520-0485(1991)021<1171:ABOAOB>2.0.CO;2.
Tsutsumi, E., T. Matsuno, R.-C. Lien, H. Nakamura, T. Senjyu, and X. Guo, 2017: Turbulent mixing within the Kuroshio in the Tokara Strait. J. Geophys. Res. Oceans, 122, 7082–7094, https://doi.org/10.1002/2017JC013049.
Waterman, S., A. C. Naveira Garabato, and K. L. Polzin, 2013: Internal waves and turbulence in the Antarctic circumpolar current. J. Phys. Oceanogr., 43, 259–282, https://doi.org/10.1175/JPO-D-11-0194.1.
Waterman, S., K. L. Polzin, A. C. Naveira Garabato, K. L. Sheen, and A. Forryan, 2014: Suppression of internal wave breaking in the Antarctic circumpolar current near topography. J. Phys. Oceanogr., 44, 1466–1492, https://doi.org/10.1175/JPO-D-12-0154.1.
Whalen, C. B., L. D. Talley, and J. A. MacKinnon, 2012: Spatial and temporal variability of global ocean mixing inferred from Argo profiles. Geophys. Res. Lett., 39, L18612, https://doi.org/10.1029/2012GL053196.
Whalen, C. B., J. A. MacKinnon, L. D. Talley, and A. F. Waterhouse, 2015: Estimating the mean diapycnal mixing using a finescale strain parameterization. J. Phys. Oceanogr., 45, 1174–1188, https://doi.org/10.1175/JPO-D-14-0167.1.
Wu, Y., E. Kunze, A. Tandon, and A. Mahadevan, 2023: Reabsorption of lee-wave energy in bottom-intensified currents. J. Phys. Oceanogr., 53, 477–491, https://doi.org/10.1175/JPO-D-22-0058.1.
Wunsch, C., and R. Ferrari, 2004: Vertical mixing, energy, and the general circulation of the oceans. Annu. Rev. Fluid Mech., 36, 281–314, https://doi.org/10.1146/annurev.fluid.36.%20050802.122121.
Zemskova, V. E., and N. Grisouard, 2021: Near-inertial dissipation due to stratified flow over abyssal topography. J. Phys. Oceanogr, 51, 2483–2504, https://doi.org/10.1175/JPO-D-21-0007.1.
Zheng, K., and M. Nikurashin, 2019: Downstream propagation and remote dissipation of internal waves in the Southern Ocean. J. Phys. Oceanogr., 49, 1873–1887, https://doi.org/10.1175/JPO-D-18-0134.1.