1. Introduction
Processes that drive enhanced mixing near the sloping seafloor have received increased attention in recent years because of their potential role in shaping water-mass transformation and diapycnal upwelling (Ferrari et al. 2016; McDougall and Ferrari 2017; Callies and Ferrari 2018). One such process is critical reflection of inertia-gravity waves (IGWs) which occurs when wave rays align with bathymetry such that upon reflection, wave energy is focused near the bottom, leading to bores, boluses, vortices, turbulence, and mixing (Cacchione and Wunsch 1974; Kunze and Llewellyn Smith 2004; Chalamalla et al. 2013). The phenomenon has almost exclusively been studied with the internal tides in mind since they carry a significant fraction of the energy in the oceanic internal wave field and because many continental slopes are near-critical for tidal frequencies (Cacchione et al. 2002). Near-inertial waves (NIWs) carry a comparable amount of energy and have a power input into them that is similar to the internal tides (Alford 2003; Ferrari and Wunsch 2009), but they have not been considered as key players in driving mixing via critical reflection on sloping topography. It seems reasonable to neglect NIWs in this regard, because according to classical internal wave theory, NIWs propagate at very shallow angles and therefore would only experience critical reflection off nearly flat bathymetry, which would not result in much wave amplification. However, classical internal wave theory does not account for the modification of wave propagation by background flows.
It is not unusual to find flows in the ocean with isopycnals that follow bathymetry. Dense overflows, such as those found on the western Weddell Sea margin, in the Denmark Strait, or over the Iceland–Faroe Ridge, for example, naturally generate bottom-intensified along-isobath currents where the isopycnals that encapsulate their dense waters blanket topographic features (Muench and Gordon 1995; Girton et al. 2001; Beaird et al. 2013). Isopycnals can also be aligned with bathymetry by upslope Ekman flows associated with the Ekman arrest of currents flowing opposite to the direction of Kelvin wave propagation (Garrett et al. 1993). The Florida Current is an example of such a flow and indeed has isopycnals that tend to parallel the continental slope off of Florida (Winkel et al. 2002). Wind-forced coastal upwelling can also result in isopycnals paralleling the bottom, and there is evidence that NIWs are amplified in critical layers during periods of upwelling but not during downwelling (Federiuk and Allen 1996).
Another example of flow that can meet the sρ = α criterion are the currents on the inshore side of the anticyclonic eddies that form in the Mississippi/Atchafalaya River plume on the Texas–Louisiana shelf. High-resolution hydrographic sections made on this shelf as part of the Mechanisms Controlling Hypoxia study (e.g., Zhang et al. 2015) illustrate the structure of the density field associated with these eddies (Fig. 1). Density surfaces form a bowl-like structure within the anticyclones while near the bottom isopycnals create a stratified layer that shoals toward the shore with sρ ≈ α.
During summer, the anticyclones on the Texas–Louisiana shelf coincide with strong near-inertial currents driven by the diurnal land–sea breeze which is near-resonant since the diurnal frequency is close to f (Zhang et al. 2009). Therefore if these near-inertial currents create downward propagating waves, then the anticyclones would provide the ideal conditions for critical reflection of NIWs over sloping topography. Realistic simulations of the circulation and wave field on the Texas–Louisiana shelf suggest that these conditions are indeed met. We describe these simulation in section 2 and use them to motivate theoretical analyses (section 3) and idealized simulations (section 4) aimed at understanding the underlying physics behind the phenomenon. With this coastal scenario as an example, the ultimate goal of this study is to build the link between wave trapping within a slantwise critical layer and the enhanced bottom mixing that it can induce. We end the article with discussions of the enhanced diapycnal transport in bottom critical layers and the mixing enhancement (section 5), which is followed by a summary of our conclusions (section 6).
2. Realistic simulations of NIW–eddy interactions on the Texas–Louisiana shelf
Here, we present results from the TXLA model, a realistic simulation on the Texas–Louisiana shelf in the northern Gulf of Mexico that highlights the interaction of NIWs with anticyclones in a coastal region with sloping bathymetry (Zhang et al. 2012). In the northern Gulf of Mexico, the Mississippi, and Atchafalaya Rivers create a large region of buoyant, relatively freshwater over the Texas–Louisiana shelf. The river plume front is unstable to baroclinic inabilities during summertime, due to a pooling of freshwater over the Louisiana shelf by weak upwelling winds and a lack of storm fronts, which generates a rich field of eddies (Hetland 2017; Qu and Hetland 2020). As illustrated in the TXLA model output, the eddies are characteristically fresh (buoyant) anticyclones, surrounded by strong cyclonic filaments at their edges (Figs. 2a,b). In addition, since storms are infrequent in the summertime and winds are generally mild, the diurnal land–sea breeze becomes an important forcing mechanism (Fig. 2c). Noting that this region is near the critical latitude, 29°N, the diurnal land–sea breeze is nearly resonant with the local inertial frequency, such that the land–sea breeze drives significant near-inertial oscillations, with peak clockwise rotating velocities of around 0.5 m s−1 (Fig. 2d). There are indications that these oscillations at the surface become downward propagating NIWs that radiate away from the offshore edge of anticyclones toward the shoaling bathymetry. Namely, bands of vertical shear in the zonal and meridional velocities descend from the offshore edge of the eddy and bend upward with isopycnals near the bottom on the inshore side of the eddy (Figs. 2h,i). The shear bands propagate upward (not shown) indicating upward phase propagation and hence suggesting downward energy propagation.
Interestingly, the turbulent kinetic energy (TKE) dissipation is enhanced near the bottom where the waves are approaching (marked by the green box in Fig. 2f) with values that are comparable to the dissipation near the surface. The TKE dissipation rate ε is diagnosed via the k–ε turbulence closure scheme. The bottom dissipation pulses over an inertial period (Fig. 2e) indicating a relationship between the enhanced dissipation and the resonantly forced near-inertial motions. We explore the underlying physics behind this relationship using theory and idealized simulations in the next two sections.
3. Theory
In this section, we develop a simple theoretical model to interpret and link the three key features revealed by the TXLA simulation: 1) downward propagation of near-inertial energy from the surface, 2) upward bending of shear bands near the bottom, and 3) enhanced dissipation in the stratified layer over the bottom. The theoretical model that integrates these key elements is schematized in Fig. 3 and is elaborated on below.
a. Downward propagation of near-inertial wave from the surface
However, to vertically propagate, NIWs need to acquire a finite horizontal wavelength and a nonzero horizontal wavenumber. The horizontal wavelength of NIWs can be reduced due to the presence of vorticity gradients, via the process of refraction (Young and Jelloul 1997; Asselin and Young 2020). Gradients in
b. Reversal of vertical energy propagation in the anomalously low-frequency regime
The upward bending of the shear bands at depth on the inshore side of the anticyclone seen in the TXLA simulation (e.g., Figs. 2h,i) suggests that the vertical propagation of the surface-generated NIWs changes sign at depths well above the bottom. Such a reversal of vertical energy propagation not due to bottom reflections is possible in background flows with baroclinicity. This follows from the expression for the slope of wave rays (1) which can change sign without switching characteristics [i.e., without switching roots in (1), which occurs at reflections]. The vertical direction of energy propagation reverses sign where sray = 0, which occurs where the wave’s frequency is equal to the local effective inertial frequency ω = feff. In a flow that is baroclinic, since ωmin < feff, the waves can propagate past this location and when they do so, sray and the vertical component of their group velocity changes sign. In this region, the wave’s frequency is less than feff but greater than ωmin; that is, ωmin < ω < feff. This is the so-called anomalously low-frequency regime defined by Mooers (1975), where NIWs are characterized by unusual behavior. In particular, the vertical components of the group and phase velocities can be in the same direction in the anomalously low-frequency regime (Whitt and Thomas 2013). This is observed in the TXLA simulation, since the shear bands that bend upward near the bottom on the inshore side of the eddy (thus fluxing energy to the shallows) also propagate upward in time, indicating a positive phase velocity.
This process is schematized in Fig. 3 for inertial waves. The location where the wave rays start to bend is where sray = 0 and feff = f. After passing the bending location, sray increases from zero so that the wave rays bend upward. At the same time, the waves enter the anomalously low-frequency regime (where feff > f), and according to the theory, the phase velocity should have the same sign as the group velocity such that the phase also propagate upward.
c. Trapping in a slantwise critical layer
Under these conditions, while the minimum frequency is inertial, ωmin = f, the effective inertial frequency
4. Idealized simulations
a. Base run
The Regional Ocean Modeling System (ROMS) is employed in this study, which is a free-surface, hydrostatic, primitive equation ocean model that uses an S-coordinate in the vertical direction (Shchepetkin and McWilliams 2005). ROMS is configured to conduct idealized simulations. The model domain represents an idealized coastal region over a continental shelf with a constant slope α = 5 × 10−4, and with the depths ranging from 5 to 118 m. The domain has an across-shore span of 226 km and an alongshore width of 4 km. The domain is set to be extremely narrow in the alongshore direction with few grid points so that the variation in the alongshore direction can be assumed to be negligibly small. The horizontal resolution is 220 m × 220 m. There are 64 layers in the vertical direction with the stretching parameters of θS = 3.0 and θB = 0.4. The alongshore boundary conditions are set to be periodic, and the offshore open boundary has a sponge layer that damps the waves propagating toward the open boundary. The Coriolis parameter is equal to the diurnal frequency; that is, f = 2π/(86 400) s−1 ≈ 7.27 × 10−5 s−1. The wind forcing is set to mimic a diurnal land–sea breeze—a rectilinear oscillating wind oriented in the across-shore direction with an amplitude of 4 × 10−2 N m−2. The simulation is run for 10 days.
The initial conditions correspond to an anticyclonic baroclinic flow with a slantwise critical layer onshore and a buoyant front offshore (Fig. 4), with parameters that are based on the realistic simulation. The critical layer has an across-shore width of LC = 50 km, and the offshore front has a width of L = 40 km. Note that there is a transition zone with a width of LT = 20 km in the middle where the horizontal buoyancy gradient linearly decreases to zero with increasing across-shore distance. The flow and density fields have no variations in the alongshore direction. The stratification is set to N2 = 3 × 10−3 s−2, a value based on the realistic simulation, and is constant across the domain. The density structure of the critical layer is determined by N2 and α, and the flow is initially in a thermal wind balance with the density field. The density structure of the offshore buoyant front is determined by the velocity structure due to the constraint of the thermal wind balance. In the horizontal direction, moving in the across-shore direction, the surface velocity at the offshore front increases from zero with a vorticity of ζ0 = −0.3f between LC + LT ≤ y ≤ LC + LT + L and decays exponentially to zero offshore and outside of this region. In the vertical direction, the velocity decays linearly to zero toward the bottom. There is no initial across-shore flow in the domain. The “MPDATA” scheme is used for the tracer advection (Smolarkiewicz and Margolin 1998). The k–ε turbulence closure scheme is used to parameterize vertical mixing, and the Canuto A stability function formulation is applied (Umlauf and Burchard 2003; Canuto et al. 2001). No explicit lateral diffusivity is used in the simulation. The parameters used to configure this simulation are listed in the first row of Table 1.
Parameters used in the base run (in the first row) and ensemble runs (in the second row), where α is bottom slope; f is Coriolis parameter; τ is the amplitude of the oscillatory, across-slope wind stress; LC, LT, and L are the length scales of the critical layer, transition zone, and offshore front, respectively; ζ0 is the surface relative vorticity of the offshore front; and N2 is the stratification in the nonrotated coordinates. Only L, ζ0, and N2 vary in the ensemble simulations, and there are a total of 18 ensemble runs.
Under the resonant wind forcing, NIWs start to develop in the first few inertial periods and then enhanced bottom mixing follows. Snapshots of the vertical shear after four inertial periods reveal the presence of shear bands (Figs. 5a,b). The orientation of the shear bands suggests that the NIWs are generated at the offshore front, where the gradient in relative vorticity is largest. This is consistent with the theoretical prediction that the horizontal wavelength of NIWs shrinks in regions with strong vorticity gradients so that the waves can propagate vertically (Young and Jelloul 1997; Asselin and Young 2020). The slantwise shear bands imply that the NIWs are propagating vertically and bend upward when approaching the bottom. Furthermore, mixing is enhanced within the bottom critical layer, which corresponds to the area where the waves are focused (Fig. 5d). This idealized simulation qualitatively reproduces the phenomena found in the realistic simulation (Fig. 2).
To understand the pattern of wave propagation suggested by the shear bands, ray tracing is conducted by applying the Wentzel–Kramers–Brillouin (WKB) approximation. The WKB approximation is only valid when the background flow field does not significantly change over the scales of the waves. We will accept this approximation a priori and then validate it below by demonstrating a consistency with an energetics analysis. The initial fields of density and velocity (Fig. 4) are used for the background flow in the ray-tracing calculation. The details of the calculation are described in appendix B. Rays are initiated at z = −2 m at the offshore end of the front with 3 km spacing. The ray paths have a similar shape to the shear bands and indicate that wave energy is radiated downward from the surface. As the waves enter the anomalously low-frequency regime (marked in the bottom panel of Fig. 4), they bend such that the slopes of wave rays are near zero. When the waves approach the critical layer, the waves slow down and eventually get trapped as |cg| → 0 (Fig. 5c). Moreover, the rays converge at the location, where the bottom mixing is enhanced, implying that wave trapping within the critical layer might be the mechanism enhancing the bottom mixing.
The mixing in the critical layer exhibits an oscillatory behavior, which is reflected in the temporal variations in the TKE dissipation rate and turbulent buoyancy flux. The TKE dissipation rate ε is diagnosed via the k–ε turbulence closure scheme. The magnitude of the the turbulent buoyancy flux is parameterized as κN2, where κ is the turbulent diffusivity also diagnosed from the k–ε closure. The mean values of ε and κN2 are calculated within the control volume (marked by the green box), and the temporal variations of these measures are shown in (Figs. 5e,f). Both ε and κN2 exhibit inertial pulsing, implying that the bottom mixing is enhanced at the inertial frequency. This reproduces the inertial pulsing of ε found in the realistic simulation (Fig. 2e) and strengthens the link between the bottom enhanced mixing and the NIWs.
b. Energetics
The energetics of the waves in the idealized simulation are analyzed, using (10), within the control volume marked by the green box in Fig. 5d, where the mixing is enhanced. Each term in (10) is integrated over the control volume to obtain time series (Fig. 6).
c. Comparative run
To demonstrate the effect of coastal fronts on vertically radiating NIWs, we present a comparative simulation without the eddy-like front to contrast the response of NIWs and bottom mixing to the simulation with the eddy-like front (i.e., the base run discussed above). The initial density field is shown in Fig. 7. The difference with the setup of the base run is the absence of the lateral buoyancy gradients, and hence also no background flow, all other parameters are the same (Table 1).
A “two layer” response is found to be dominant in the comparative simulation; the surface and bottom velocities are out of phase by nearly 180 degrees (Figs. 7e,f). Such a two-layer structure has been observed in many coastal seas (Orlić 1987; van Haren et al. 1999; Knight et al. 2002; Rippeth et al. 2002) and also some large lakes (Malone 1968; Smith 1972). This response is attributed to the presence of a coastal boundary (Davies and Xing 2002). The presence of the coastal boundary yields a pressure gradient at depth, which drives the inertial current in the lower layer and leads to a 180° phase shift with the upper layer (Xing and Davies 2004). Consistent with the observations, the comparative run shows that the first baroclinic mode dominates the response in a coastal system without fronts or currents.
The differences between the comparative and base runs are summarized as follows. First, no clear slantwise shear bands exist in the interior (Figs. 7c,d), suggesting that there is no significant vertical radiation of NIWs from the surface. Second, the dissipation does not exhibit clear inertial pulsing that has been observed in the base run (Fig. 7g); it pulses semi-inertially as the bottom velocity attains peaks. Last, the bottom boundary layer over the shelf is thin (Fig. 7b), and the response of the dissipation is much weaker than that in the base run (Figs. 7e,g). Overall, the comparative simulation indicates that the eddy-like coastal front is essential for the vertical radiation of NIWs and therefore the bottom enhanced mixing.
d. Parameter dependence of wave trapping and mixing in the critical layer
The idealized simulations are used to explore the dependence of wave trapping and mixing in the critical layer in the framework of the idealized configuration. There are 8 controlling parameters and they are listed in Table 1. To efficiently explore this parameter space, we fix the external parameters (i.e., the background rotation f, wind forcing τ, and bottom slope α) as well as the dimension of the nearshore front (i.e., the length scales of the critical layer and transition zone, LC and LT), but vary the parameters associated with the offshore front (i.e., the frontal width L, relative vorticity ζ0, and stratification N2). In other words, we fix the properties of the critical layer but vary the parameters that influence the propagation of waves toward the critical layer. By varying these parameters we can quantify the sensitivity of the dissipation and mixing to the degree of wave trapping. A total of 18 simulations (including the base run) were performed and are listed in the second row of Table 1.
In concert with the ROMS simulations, ray tracing is conducted for each run. Rays are initiated at z = −2 m within the offshore front across the width L, separated by a spacing of dy = 1 km, and traced according to the procedure described in appendix B. Rays either reach the critical layer or hit the bottom and reflect offshore. To quantify the wave trapping in the critical layer from the ray-tracing solutions, we define the trapping ratio γ, the ratio between the number of the rays reaching the critical layer and the total number of the rays. A higher value of γ indicates a larger portion of wave energy that reaches the trapping zone and hence represents a highly trapped scenario. The trapping ratio is a metric that concisely captures the parameter dependence of wave trapping in the critical layer.
Relative vorticity modifies the minimum frequency ωmin, such that stronger anticyclonic vorticity allows NIWs to propagate more vertically, e.g., (1). A subset of the ensemble simulations run with different values of the vorticity ζ0 but with a fixed frontal width of L = 40 km, and stratification of N2 = 5 × 10−3 s−2 illustrates this physics (Fig. 8). The stronger the anticyclonic vorticity, the more steep the rays are, and they miss the critical layer. For instance, in the case with ζ0 = −0.7f, there are fewer rays reaching the critical layer (denoted by red) and more rays reflect offshore (denoted by green) than in the other two runs with weaker vorticity. Correspondingly, the case with ζ0 = −0.7f has the lowest TKE dissipation rate, suggesting that reduced wave trapping leads to weaker mixing.
The dimension of the offshore front also modulates the wave trapping. This is illustrated in Fig. 9 for a group of ensemble runs with various frontal widths L but with fixed relative vorticity (i.e., ζ0 = −0.5f) and stratification (N2 = 3 × 10−3 s−2). Noting that the difference in vorticity across the jet is the same for all three cases, the propagation of NIWs only depends on the geometry of the offshore front. The ray-tracing solutions shown in Fig. 9 demonstrate that the wider fronts have fewer rays that reach the critical layer and get trapped. This is because wider fronts move the rays away from the critical layer. Consquently, the run with L = 50 km has the weakest TKE dissipation rate and mixing.
Last, we quantify the relation between wave trapping and mixing using the trapping ratio γ. First, to test the skill of the parameter γ in predicting the degree of wave trapping, we calculate the maximum (in time, over one inertial period) volume-integrated WEF
5. Discussion
a. Enhanced diapycnal transport in the critical layer
Another way to quantify the diapycnal transport is to track the diapycnal movement of a passive tracer. To this end, a passive tracer was released in both the base run and the comparative run to contrast the bottom diapycnal transport in simulations with and without wave trapping. The tracer is initialized in the first four sigma layers above the bottom with a concentration equal to one (Figs. 13a,b). The concentration outside of this layer is set to zero. The tracer is released at t = 90 h and monitored for three inertial periods. In terms of the spatial distribution, at t = 106 h (the time of the peak wd), the base run shows a significant reduction of the tracer at the location where the diapycnal velocity is enhanced and the waves are trapped (Fig. 13c). In terms of temporal variability, in the region with enhanced wd, the variation of the tracer concentration suggests that the tracer is transported out of the bottom layer during the period (from 103 to 109 h) when the diapycnal velocity is enhanced (Fig. 13e). In contrast, the comparative run does not show such a significant reduction in the tracer concentration near the bottom (Figs. 13d,e).
The distribution function calculated from the base run indicates that the tracer migrates to lighter density classes with time (see the top panel of Fig. 14). When the diapycnal velocity is largest (at t = 106 h), there is a convergence of the tracer toward a narrow density class. This highlights the role of the diapycnal velocity in transporting the tracer. Also, the convergence can be seen by contrasting ∂M/∂ρ at the time when wd is maximum with the one at the initial time (see the bottom panel of Fig. 14). Furthermore, the convergence of the tracer persists with time, confirming that the enhanced diapycnal velocity does effectively transport the tracer across isopycnals.
The enhancement of the diapycnal transport by wave trapping has implications for coastal biogeochemistry and ecosystems. In coastal zones, freshwater from rivers strengthens the stratification and can suppresses the ventilation of bottom waters. This combined with phytoplankton blooms fueled by nutrients in the freshwater can lead to bottom hypoxia and the formation of “dead zones” (Bianchi et al. 2010). One region where bottom hypoxia often occurs is the Texas–Louisiana shelf where the development of the hypoxia is heavily modulated by near-inertial motions (Xomchuk et al. 2020). We have demonstrated that NIW trapping within critical layers is potentially active, suggesting that mixing of the stratified bottom waters by this process could potentially ventilate these oxygen poor waters. In fact, intrusions of hypoxic waters emanating from slantwise stratified layers near the bottom have been observed on the shelf during the MCH survey suggesting active mixing in these layers (Zhang et al. 2015).
b. The inertial pulsing of mixing
The influence of WEF in shaping the evolution of the vertical structure of the shear, stratification, dissipation, and mixing is shown in (Fig. 15). As the wave energy flux converges (e.g., Figure 6 between hours 99 and 105), the vertical shear and DKE increases (Figs. 15b,d). Consequently, ε and κN2 are enhanced via the k–ε model (Figs. 15e,f) and the stratification is reduced (Fig. 15a). At this stage, the total Richardson number Ri decreases to around 0.25, indicating that the criterion for shear instabilities is crossed (Fig. 15c).
It might seem counterintuitive that the wave energy flux pulses inertially rather than semi-inertially. Based on the polarization relation for linear NIWs, the wave pressure and across-front velocity are in phase so that the wave energy flux should exhibit a semi-inertial response (Whitt and Thomas 2013). This theoretical prediction breaks down in the simulations because the waves are of finite amplitude and hence their dynamics is nonlinear.
To highlight the nonlinear wave dynamics, the base run from the idealized simulations (discussed in section 4a) is rerun with a wind stress that is one order of magnitude weaker (4 × 10−3 N m−2) and the two solutions are contrasted. Both simulations have inertial oscillations in the across-shore velocity which vary fairly symmetrically over a wave period (Fig. 16a). In contrast, the across-shore pressure gradient force (PGF) in the original base run exhibits an asymmetric oscillation, as opposed to the regular inertial oscillations in the PGF in the weak-wind run (Fig. 16b). As a consequence the convergence of wave energy flux does not oscillate at twice the inertial frequency unlike in the weak-wind run (Fig. 16c). This suggests that the PGF is asymmetrically modified by the finite amplitude of the waves, and this is what causes the convergence of wave energy flux to pulse inertially rather than semi-inertially.
6. Conclusions
A specific type of NIW critical layer over sloping bathymetry is explored in this study. When isopycnals align with sloping bathymetry in a stratified layer, a critical layer for NIWs with ω = f forms. Upon entering this critical layer, the waves are trapped and amplified since their group velocity goes to zero, and mixing is enhanced.
Such slantwise critical layers form in a fully three-dimensional realistic simulation of anticyclonic eddies on the Texas–Louisiana shelf. The realistic simulation exhibits an inertial enhancement of bottom mixing where the energy from surface-generated NIWs is focused in bottom stratified layers on the shelf. Idealized, two-dimensional ROMS simulations reproduce these phenomena, and ray tracing and analyses of the waves’ energetics support the idea that the enhanced bottom mixing is caused by the convergence of NIW energy in slantwise critical layers and largely follows this two-dimensional physics. This conclusion is based on results from an ensemble of simulations that cover the relevant parameter space. The ensemble runs show that background flows that more effectively trap wave rays result in stronger wave energy convergence in the critical layer and enhanced mixing. Overall, the link between enhanced mixing and wave trapping is motivated by the realistic simulation, understood using the theoretical analyses, and strengthened by the results from the idealized simulations and ray-tracing solutions.
Although the focus of this study is a particular application on the Texas–Louisiana shelf, the mechanism of NIW amplification in critical layers over sloping bathymetry should be active in other settings. For example, another coastal application could be upwelling systems over continental shelves, where upwelled, dense waters blanket bathymetry. Potential examples include the upwelling systems over the Oregon shelf (Federiuk and Allen 1996; Avicola et al. 2007), the New Jersey inner shelf (Chant 2001), the shelf off of the California coast (Nam and Send 2013; Woodson et al. 2007), and the Tasmanian shelf, where recent observations suggest evidence of enhanced near-inertial energy and wave trapping in slantwise critical layers (Schlosser et al. 2019). In these upwelling systems, Federiuk and Allen (1996) highlight the importance of background flows in modifying the group velocity of NIWs and attribute the observed enhancement of near-inertial energy during periods of upwelling versus downwelling to wave trapping, similar to the mechanism that we have described here. However Federiuk and Allen (1996) did not identified the key criterion for NIW critical layer formation–alignment of isopycnals with bathymetry–that we have determined from our analyses.
Examples of open-ocean flows that can form such critical layers include dense overflows and currents that drive upslope Ekman arrest in bottom boundary layers. One example of the latter is the Kuroshio Current flowing over the continental slope southeast of Kyushu, Japan (Nagai et al. 2019). Over the shelfbreak, isopycnals tend to align with the bottom suggesting the existence of a slantwise critical layer. In this layer slantwise shear bands suggestive of amplified NIWs are found and coincide with regions where the turbulent dissipation rate is elevated to values of O(10−7) m2 s−3. Another example is the Florida Current on the western side of the Straits of Florida. On this side of the straits, the alignment between the isopycnals and the continental slope indicates a slantwise critical layer, where in fact observations show that turbulence can be enhanced in stratified layers off the bottom (Winkel et al. 2002). We plan to study the dynamics of these open-ocean NIW critical layers in future work.
Diapycnal transport within these critical layers can also be enhanced due to turbulence driven by wave trapping. Such diapycnal transport can influence the distribution of biogeochemical tracers such as iron and oxygen and thus potentially influence coastal ecosystems. In the open ocean, NIW trapping in critical layers could affect abyssal diapycnal transport near the bottom, which could modify water-mass distributions and influence the meridional overturning circulation.
Acknowledgments
This work was funded by the SUNRISE project, NSF Grants OCE-1851450 (to authors Qu and Thomas) and OCE-1851470 (to author Hetland). We thank Olivier Asselin, Bertrand Delorme, Jinliang Liu, Jen MacKinnon, Jonathan Nash, Guillaume Roullet, Kipp Shearman, and John Taylor for very helpful suggestions while we prepared this paper.
APPENDIX A
Rotated and Nonrotated Coordinates
APPENDIX B
Ray Tracing
REFERENCES
Alford, M., 2003: Redistribution of energy available for ocean mixing by long-range propagation of internal waves. Nature, 423, 159–162, https://doi.org/10.1038/nature01628.
Asselin, O., and W. R. Young, 2020: Penetration of wind-generated near-inertial waves into a turbulent ocean. J. Phys. Oceanogr., 50, 1699–1716, https://doi.org/10.1175/JPO-D-19-0319.1.
Avicola, G., J. N. Moum, A. Perlin, and M. D. Levine, 2007: Enhanced turbulence due to the superposition of internal gravity waves and a coastal upwelling jet. J. Geophys. Res., 112, C06024, https://doi.org/10.1029/2006JC003831.
Beaird, N. L., P. Rhines, and C. Eriksen, 2013: Overflow waters at the Iceland–Faroe Ridge observed in multiyear Seaglider surveys. J. Phys. Oceanogr., 43, 2334–2351, https://doi.org/10.1175/JPO-D-13-029.1.
Bennett, S., 1986: The relationship between vertical, diapycnal, and isopycnal velocity and mixing in the ocean general circulation. J. Phys. Oceanogr., 16, 167–174, https://doi.org/10.1175/1520-0485(1986)016<0167:TRBVDA>2.0.CO;2.
Bianchi, T. S., S. DiMarco, J. Cowan Jr., R. Hetland, P. Chapman, J. Day, and M. Allison, 2010: The science of hypoxia in the northern Gulf of Mexico: A review. Sci. Total Environ., 408, 1471–1484, https://doi.org/10.1016/j.scitotenv.2009.11.047.
Cacchione, D., and C. Wunsch, 1974: Experimental study of internal waves over a slope. J. Fluid Mech., 66, 223–239, https://doi.org/10.1017/S0022112074000164.
Cacchione, D., L. Pratson, and A. Ogston, 2002: The shaping of continental slopes by internal tides. Science, 296, 724–727, https://doi.org/10.1126/science.1069803.
Callies, J., and R. Ferrari, 2018: Dynamics of an abyssal circulation driven by bottom-intensified mixing on slopes. J. Phys. Oceanogr., 48, 1257–1282, https://doi.org/10.1175/JPO-D-17-0125.1.
Canuto, V. M., A. Howard, Y. Cheng, and M. Dubovikov, 2001: Ocean turbulence. Part I: One-point closure model—Momentum and heat vertical diffusivities. J. Phys. Oceanogr., 31, 1413–1426, https://doi.org/10.1175/1520-0485(2001)031<1413:OTPIOP>2.0.CO;2.
Chalamalla, V., B. Gayen, A. Scotti, and S. Sarkar, 2013: Turbulence during the reflection of internal gravity waves at critical and near-critical slopes. J. Fluid Mech., 729, 47–68, https://doi.org/10.1017/jfm.2013.240.
Chant, R. J., 2001: Evolution of near-inertial waves during an upwelling event on the New Jersey inner shelf. J. Phys. Oceanogr., 31, 746–764, https://doi.org/10.1175/1520-0485(2001)031<0746:EONIWD>2.0.CO;2.
Davies, A. M., and J. Xing, 2002: Influence of coastal fronts on near-inertial internal waves. Geophys. Res. Lett., 29, 2114, https://doi.org/10.1029/2002GL015904.
Federiuk, J., and J. Allen, 1996: Model studies of near-inertial waves in flow over the Oregon continental shelf. J. Phys. Oceanogr., 26, 2053–2075, https://doi.org/10.1175/1520-0485(1996)026<2053:MSONIW>2.0.CO;2.
Ferrari, R., and C. Wunsch, 2009: Ocean circulation kinetic energy: Reservoirs, sources, and sinks. Annu. Rev. Fluid Mech., 41, 253–282, https://doi.org/10.1146/annurev.fluid.40.111406.102139.
Ferrari, R., A. Mashayek, T. McDougall, M. Nikurashin, and J. Campin, 2016: Turning ocean mixing upside down. J. Phys. Oceanogr., 46, 2239–2261, https://doi.org/10.1175/JPO-D-15-0244.1.
Garrett, C., P. MacCready, and P. Rhines, 1993: Boundary mixing and arrested Ekman layers: Rotating stratified flow near a sloping boundary. Annu. Rev. Fluid Mech., 25, 291–323, https://doi.org/10.1146/annurev.fl.25.010193.001451.
Girton, J. B., T. B. Sanford, and R. H. Käse, 2001: Synoptic sections of the Denmark Strait overflow. Geophys. Res. Lett., 28, 1619–1622, https://doi.org/10.1029/2000GL011970.
Hetland, R. D., 2017: Suppression of baroclinic instabilities in buoyancy-driven flow over sloping bathymetry. J. Phys. Oceanogr., 47, 49–68, https://doi.org/10.1175/JPO-D-15-0240.1.
Knight, P., M. Howarth, and T. Rippeth, 2002: Inertial currents in the northern North Sea. J. Sea Res., 47, 269–284, https://doi.org/10.1016/S1385-1101(02)00122-3.
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<0544:NIWPIG>2.0.CO;2.
Kunze, E., and S. Llewellyn Smith, 2004: The role of small-scale topography in turbulent mixing of the global ocean. Oceanography, 17, 55–64, https://doi.org/10.5670/oceanog.2004.67.
Lee, D.-K., and P. P. Niiler, 1998: The inertial chimney: The near-inertial energy drainage from the ocean surface to the deep layer. J. Geophys. Res., 103, 7579–7591, https://doi.org/10.1029/97JC03200.
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<0005:SBBLOA>2.0.CO;2.
Malone, F. D., 1968: An analysis of current measurements in Lake Michigan. J. Geophys. Res., 73, 7065–7081, https://doi.org/10.1029/JB073i022p07065.
McDougall, T., and R. Ferrari, 2017: Abyssal upwelling and downwelling driven by near-boundary mixing. J. Phys. Oceanogr., 47, 261–283, https://doi.org/10.1175/JPO-D-16-0082.1.
Mooers, C. N., 1975: Several effects of a baroclinic current on the cross-stream propagation of inertial-internal waves. Geophys. Astrophys. Fluid Dyn., 6, 245–275, https://doi.org/10.1080/03091927509365797.
Muench, R. D., and A. L. Gordon, 1995: Circulation and transport of water along the western Weddell Sea margin. J. Geophys. Res., 100, 18 503–18 515, https://doi.org/10.1029/95JC00965.
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.
Nam, S., and U. Send, 2013: Resonant diurnal oscillations and mean alongshore flows driven by sea/land breeze forcing in the coastal southern California bight. J. Phys. Oceanogr., 43, 616–630, https://doi.org/10.1175/JPO-D-11-0148.1.
Orlić, M., 1987: Oscillations of the inertia period on the Adriatic sea shelf. Cont. Shelf Res., 7, 577–598, https://doi.org/10.1016/0278-4343(87)90024-0.
Pedlosky, J., Ed., 1987: Ageostrophic motion. Geophysical Fluid Dynamics, Springer, 624–669.
Qu, L., and R. Hetland, 2020: Non-geostrophic baroclinic instability over sloping bathymetry: Buoyant flow regime. J. Phys. Oceanogr., 50, 1937–1956, https://doi.org/10.1175/JPO-D-19-0145.1.
Rippeth, T. P., J. H. Simpson, R. J. Player, and M. Garcia, 2002: Current oscillations in the diurnal–inertial band on the Catalonian shelf in spring. Cont. Shelf Res., 22, 247–265, https://doi.org/10.1016/S0278-4343(01)00056-5.
Schlosser, T. L., N. L. Jones, C. E. Bluteau, M. H. Alford, G. N. Ivey, and A. J. Lucas, 2019: Generation and propagation of near-inertial waves in a baroclinic current on the Tasmanian shelf. J. Phys. Oceanogr., 49, 2653–2667, https://doi.org/10.1175/JPO-D-18-0208.1.
Shchepetkin, A. F., and J. C. McWilliams, 2005: The regional oceanic modeling system (ROMS): A split-explicit, free-surface, topography-following-coordinate oceanic model. Ocean Modell., 9, 347–404, https://doi.org/10.1016/j.ocemod.2004.08.002.
Smith, N., 1972: Temporal variations of nearshore inertial motion in Lake Superior. Proc. 15th Conf. Great Lakes Research, Madison, WI, IAGLR, 673–679.
Smolarkiewicz, P. K., and L. G. Margolin, 1998: MPDATA: A finite-difference solver for geophysical flows. J. Comput. Phys., 140, 459–480, https://doi.org/10.1006/jcph.1998.5901.
Umlauf, L., and H. Burchard, 2003: A generic length-scale equation for geophysical turbulence models. J. Mar. Res., 61, 235–265, https://doi.org/10.1357/002224003322005087.
van Haren, H., L. Maas, J. Zimmerman, H. Ridderinkhof, and H. Malschaert, 1999: Strong inertial currents and marginal internal wave stability in the central North Sea. Geophys. Res. Lett., 26, 2993–2996, https://doi.org/10.1029/1999GL002352.
van Meurs, P., 1998: Interactions between near-inertial mixed layer currents and the mesoscale: The importance of spatial variabilities in the vorticity field. J. Phys. Oceanogr., 28, 1363–1388, https://doi.org/10.1175/1520-0485(1998)028<1363:IBNIML>2.0.CO;2.
Warner, J. C., C. R. Sherwood, H. G. Arango, and R. P. Signell, 2005: Performance of four turbulence closure models implemented using a generic length scale method. Ocean Modell., 8, 81–113, https://doi.org/10.1016/j.ocemod.2003.12.003.
Wenegrat, J. O., J. Callies, and L. N. Thomas, 2018: Submesoscale baroclinic instability in the bottom boundary layer. J. Phys. Oceanogr., 48, 2571–2592, https://doi.org/10.1175/JPO-D-17-0264.1.
Whitt, D. B., and L. N. Thomas, 2013: Near-inertial waves in strongly baroclinic currents. J. Phys. Oceanogr., 43, 706–725, https://doi.org/10.1175/JPO-D-12-0132.1.
Winkel, D. P., M. C. Gregg, and T. B. Sanford, 2002: Patterns of shear and turbulence across the Florida Current. J. Phys. Oceanogr., 32, 3269–3285, https://doi.org/10.1175/1520-0485(2002)032<3269:POSATA>2.0.CO;2.
Woodson, C., and Coauthors, 2007: Local diurnal upwelling driven by sea breezes in northern Monterey Bay. Cont. Shelf Res., 27, 2289–2302, https://doi.org/10.1016/j.csr.2007.05.014.
Xing, J., and A. M. Davies, 2004: On the influence of a surface coastal front on near-inertial wind-induced internal wave generation. J. Geophys. Res., 109, C01023, https://doi.org/10.1029/2003JC001794.
Xomchuk, V. R., R. D. Hetland, and L. Qu, 2020: Small-scale variability of bottom oxygen in the northern Gulf of Mexico. J. Geophys. Res. Oceans, 126, e2020JC016279, https://doi.org/10.1029/2020JC016279.
Young, W., and M. B. Jelloul, 1997: Propagation of near-inertial oscillations through a geostrophic flow. J. Mar. Res., 55, 735–766, https://doi.org/10.1357/0022240973224283.
Zhang, W., R. D. Hetland, S. F. DiMarco, and K. Fennel, 2015: Processes controlling mid-water column oxygen minima over the Texas–Louisiana shelf. J. Geophys. Res. Oceans, 120, 2800–2812, https://doi.org/10.1002/2014JC010568.
Zhang, X., S. F. DiMarco, D. C. Smith IV, M. K. Howard, A. E. Jochens, and R. D. Hetland, 2009: Near-resonant ocean response to sea breeze on a stratified continental shelf. J. Phys. Oceanogr., 39, 2137–2155, https://doi.org/10.1175/2009JPO4054.1.
Zhang, X., M. Marta-Almeida, and R. Hetland, 2012: A high-resolution pre-operational forecast model of circulation on the Texas–Louisiana continental shelf and slope. J. Oper. Oceanogr., 5, 19–34, https://doi.org/10.1080/1755876X.2012.11020129.