1. Introduction
Internal waves carry a substantial fraction of the kinetic energy in the ocean and have small vertical scales with large vertical shears (Ferrari and Wunsch 2009). Consequently they are thought to play a crucial role in vertical mixing (e.g., Munk and Wunsch 1998; Wunsch and Ferrari 2004; Ferrari and Wunsch 2009). Global maps of ocean mixing inferred from Argo profiles using a finescale parameterization for mixing by internal waves reveal that mixing is enhanced in regions of high eddy kinetic energy, particularly near separated western boundary currents (Whalen et al. 2012). Separated western boundary currents underlie the atmospheric storm tracks where time-variable winds efficiently inject kinetic energy into the near-inertial band of the internal wave spectrum (Alford 2003). This suggests that near-inertial waves (NIWs) could facilitate the enhanced mixing. The analysis of the Argo data by Whalen et al. (2012) provides further evidence of this, showing that mixing varies in concert with the seasonal changes in near-inertial energy in the mixed layer. However, how NIWs ultimately lead to enhanced turbulence is not well understood. One proposed mechanism is through the interaction of the waves with the circulation and its mesoscale eddies.
NIWs are known to interact strongly with mesoscale features because the vorticity of balanced flows significantly modifies their frequency (Mooers 1975). A variety of observations have been made that illustrate this interaction (e.g., Perkins 1976; Weller 1982; Kunze and Sanford 1984; Kunze et al. 1995). Kunze et al. (1995), in particular, show regions of high energy dissipation in the thermocline associated with anticyclonic vorticity, which lowers the frequency of the waves and allows for the existence of trapped subinertial waves with amplified energy density, consistent with the well-established theoretical predictions of, for example, Kunze (1985), Klein and Hua (1988), Klein and Treguier (1995), Young and Ben-Jelloul (1997), Balmforth et al. (1998), and Plougonven and Zeitlin (2005).
More recent observations collected in the Gulf Stream under wintertime forcing as part of the Clivar Mode Water Dynamics Experiment (CLIMODE) reveal strong NIWs with subinertial frequencies (Marshall et al. 2009). The NIWs were characterized by horizontally coherent structures in the ageostrophic shear that were aligned with the slanted isopycnals of the pycnocline (e.g., Fig. 1) and that were coincident with regions of enhanced dissipation (Inoue et al. 2010). The banded structure in ageostrophic shear was found in the north wall of the Gulf Stream where the balanced flow is cyclonic and strongly baroclinic, that is, where the relative vorticity, horizontal density gradient, and thermal wind shear are large. Similar observations of coherent wavelike shear bands along isopycnals have also been made in other strong frontal regions by, for example, Shcherbina et al. (2003) and Rainville and Pinkel (2004) but these observations were interpreted as waves constrained by the variation of relative vorticity. The effects of baroclinicity, which lead to amplification parallel to isopycnals, were not investigated and may have helped to explain some unexpected results in these observations.
(top) The Rossby number and base-10 logarithm of the Richardson number, Rog and log(Rig), in the Gulf Stream from a section taken during the winter as part of the CLIMODE experiment. (bottom) The ageostrophic shear from the same section with potential density (black contours) superimposed at increments of 0.2 kg m−3. Ageostrophic shears were calculated by assuming the cross-stream shear ∂υa/∂z was all ageostrophic and the along-stream shear was equivalent to the total shear minus the thermal wind shear, ∂ua/∂z = ∂u/∂z + f−1∂b/∂y. They show the coherent phase structure typical of a downward propagating internal wave, yet with slopes that run parallel to isopycnals. These shears coincide with regions of low Rig suggesting that baroclinic effects may be important in the wave dynamics. The wave shear occupies a relatively small part of the section, that is, about 30 km, so the velocities are not backrotated since they were observed over just several hours. These observations are from survey S1 in Inoue et al. (2010).
Citation: Journal of Physical Oceanography 43, 4; 10.1175/JPO-D-12-0132.1
Two-dimensional primitive equation numerical simulations have also shown evidence of NIWs modified by the effects of baroclinicity in both a surface front and a coastal upwelling zone (Wang 1991; Federiuk and Allen 1996, respectively). However, the physical phenomena were not fully explored.
This body of recent observations and simulations suggests that the strong baroclinicity of frontal jets, not solely the vorticity, influences the dynamics of the NIWs that are propagating in them and can facilitate wave trapping, amplification, and breaking.1

2. Governing equations




3. Basic wave properties
In a bounded domain with constant background properties, (14) has constant coefficients and can be converted into a Sturm–Liouville eigenvalue problem. In fact, even if N varies with depth, this procedure can still be followed (e.g., Gerkema and Shrira 2005a,b). However, we are interested in the case where all the coefficients in (14) vary in two spatial dimensions. We will ultimately have to obtain a numerical solution to this problem. However, we can glean physical insights from this equation without too many further assumptions about the solution. In particular, it can be shown that two fundamental properties of near-inertial waves (NIWs)—their minimum frequency and direction of propagation—are significantly modified by baroclinicity. However, first we introduce an idealized background flow in which we illustrate the various wave phenomena throughout the paper.
a. Idealized background flow
The geostrophic velocity (shading) and potential density (contoured at 0.2 kg m−3 increments) for the idealized background flow given by (a) Eq. (16) and (b) the observations from the Gulf Stream.
Citation: Journal of Physical Oceanography 43, 4; 10.1175/JPO-D-12-0132.1
A flow with parameters similar to the wintertime Gulf Stream can be obtained from (16) by choosing (C1, n, C2, Ly, Lz) = (0.9 kg m−3, 2.5, −0.4 kg m−3, 120 km, 400 m) in (16) and letting f = 10−4 s−1. The Rossby number and Richardson number of this idealized geostrophic flow are displayed in Fig. 3 for comparison with the section crossing the Gulf Stream obtained during CLIMODE (in Fig. 1). The Gulf Stream observations are described in, for example, Marshall et al. (2009) and Inoue et al. (2010). With this choice of the tuning constants we have obtained a baroclinic flow that is relatively realistic, simple, and symmetric. It has slowly varying properties, but contains regions where Rog and Rig are O(1) similar to the observations.
(top) log10 (Rig) and (bottom) Rog for the idealized background flow (16). Potential density contours with an increment of 0.2 kg m−3 are superimposed. These parameters are O(1) in the idealized jet similarly to the winter Gulf Stream as shown in Fig. 1.
Citation: Journal of Physical Oceanography 43, 4; 10.1175/JPO-D-12-0132.1
b. Minimum frequency




The frequency of oscillation of fluid parcels displaced at an angle θδ with respect to the horizontal, for a basic state without a background flow (Rog = 0, Rig = ∞, dashed), with a barotropic flow (Rog = −
Citation: Journal of Physical Oceanography 43, 4; 10.1175/JPO-D-12-0132.1
Rays, which run parallel to characteristics, illustrate the downward propagation of a wave packet with a frequency (a),(c) 0.95f and (b),(d) 0.98f in both (left) the idealized flow and (right) the Gulf Stream. Baroclinic effects expand the region where rays are confined from ω ≥ F (inside the green contour) to ω ≥ ωmin (inside the magenta contour) and change the angle at which rays reflect from horizontal to the isopycnal angle. Note that there are regions where the contour marking ω = ωmin runs parallel to isopycnals. In these regions, the conditions for a slantwise critical layer are satisfied: multiple rays converge, the group velocity goes to zero, and waves are trapped. In the Gulf Stream observations, these slantwise critical layers are coincident with the region of high observed ageostrophic shear, as shown in Fig. 1.
Citation: Journal of Physical Oceanography 43, 4; 10.1175/JPO-D-12-0132.1
Schematic illustrating the physics of the minimum frequency for NIWs in geostrophic flows. Lines of constant geostrophic absolute momentum, Mg = ug − fy (solid gray), and constant buoyancy bg (dashed black) are plotted in each frame. For (a) a flow with no relative vorticity or baroclinicity, (b) a barotropic flow with anticyclonic relative vorticity, (c) a baroclinic flow with only a vertical thermal wind shear, and (d) a flow with zero PV. The minimum frequency occurs when oscillations are along lines of constant buoyancy. When the gradient of Mg along bg is weaker than that shown in (a), the restoring force is reduced and the frequency is lower than f. In (d) this gradient is zero, hence there is no restoring force and ωmin = 0.
Citation: Journal of Physical Oceanography 43, 4; 10.1175/JPO-D-12-0132.1
Before moving on, we highlight the important observation that (19) admits solutions where ωmin < F. In fact, for S2 ≠ 0, ωmin is always less than F. As discussed in Mooers (1975) and in more detail here, near-inertial waves have particularly unusual and counterintuitive behavior when ω < F. Therefore, it is useful to analyze the physics of NIWs in two separate frequency bands: the classical frequency range where ω > F and a range where ω < F that Mooers (1975) referred to as the anomalously low frequency band. Whether a wave is in one region or the other depends on its particular frequency and the background geostrophic flow that determines F and ωmin. As mentioned above, the separation between these two regions in both our idealized flow and the section from the Gulf Stream is illustrated in Fig. 5. The anomalously low frequency region is delineated by the contours where ω = F (green) and ω = ωmin (magenta) and is found where isopycnals steepen, the flow becomes more strongly baroclinic (i.e., Rig decreases), and the vertical vorticity and Rog increase.
c. Characteristics
The separation between the classical and anomalously low frequency regions for a wave with frequency ω = 0.95 in our idealized background flow and the slopes of the wave’s characteristics (top) λ+ and (bottom) λ−. The checkerboard pattern indicates that the parcel oscillations along that characteristic are unstable with respect buoyancy. The wavy line pattern indicates that parcel oscillations are unstable with respect to momentum. The green contour is where ω = F; the magenta where ω = ωmin. The basic physics of parcel oscillations in these two quasi-unstable regimes is illustrated in Fig. 9.
Citation: Journal of Physical Oceanography 43, 4; 10.1175/JPO-D-12-0132.1
Schematic of characteristics ξ+ and ξ−: There are three regimes depending on the frequency (left) above the effective inertial frequency F, (middle) at the effective inertial frequency, and (right) below the effective inertial frequency. Waves with ω < F are defined to exist in the anomalously low frequency range. If the sign of the isopycnal slope is switched, that is, θb < 0, the signs of the characteristic slopes are switched as shown in Fig. 7.
Citation: Journal of Physical Oceanography 43, 4; 10.1175/JPO-D-12-0132.1
d. Physical interpretation
To obtain a physical interpretation of these results, we now examine how an infinitesimal parcel of fluid would oscillate if it were given an instantaneous small ageostrophic velocity in a baroclinic background flow.



Schematics of parcel displacements in a mean flow with baroclinicity and force diagrams illustrating how the components of the Coriolis force,
Citation: Journal of Physical Oceanography 43, 4; 10.1175/JPO-D-12-0132.1
Parcel arguments can also be used to understand the physics at the minimum frequency, illustrated in Fig. 6. It follows from (30) that the angle that minimizes the frequency corresponds to parcel displacements that run parallel to isopycnals, that, θδ = θb ≈ S2/N2. Consequently, for this frequency the buoyancy force (28) is zero and plays no role in dynamics of the oscillation. The restoring force of these oscillations arises solely from the Coriolis force and conservation of absolute momentum (27). The strength of this restoring force, and hence the frequency that it induces, is proportional to the change in geostrophic absolute momentum that a fluid parcel experiences as it moves along a density surface. The change in Mg thus scales with the along-isopycnal gradient in Mg, which is proportional to the PV, hence explaining the formulas for ωmin, (19) and (21).
As mentioned above, Fig. 4 illustrates this dispersion relation (30) for three different background conditions: one where there is no variation in the flow (F2 = f2 and S2 = 0, the classic case), one where there is only a horizontal gradient in momentum (i.e., ζg = −0.7f, S2 = 0 and the flow is barotropic), and one where baroclinicity is important and there is a wide anomalously low frequency range (F2 = f2 and S2 = 0.03N2). In the first and second cases, S2 = 0 and the minimum frequency coincides with purely horizontal displacements (θδ = 0) and all other frequencies correspond to two parcel displacements with equal but opposite angles. In the baroclinic case in contrast, where frequencies below F are anomalously low, the angle of parcel displacement at the minimum frequency is significantly different from horizontal; it is equal to the isopycnal slope.
What is also unusual is that for frequencies ωmin < ω < F the two parcel displacements that result in a given wave period have angles
For the shallower characteristic, corresponding to parcel displacements at an angle
On the steeper characteristic, corresponding to parcel displacements at an angle
In certain regions of frequency space (occurring both when ω < F and ω > F but certainly not for all ω), the angle of parcel displacements on the steeper characteristic can be larger than the slope of Mg surfaces (i.e.,
The imaginary part of the quantity ua/υa, given by (32), for a wave with ω = 0.95f in the idealized background flow for the two characteristics: (a) ξ+ and (b) ξ−. The magnitude of this quantity indicates the local ellipticity of the hodograph, whereas the sign indicates the direction of rotation with either depth or frequency. Not only does the ellipticity differ for the two characteristics but the sense of rotation switches sign when parcel oscillations are unstable to momentum, that is, steeper than geostrophic absolute momentum surfaces (see Fig. 7). (c) The polarization relation at ωmin as a function of Rog and
Citation: Journal of Physical Oceanography 43, 4; 10.1175/JPO-D-12-0132.1



4. Energetics and propagation
In this section, we will characterize how near-inertial waves transport energy when propagating across baroclinic currents. To address this issue, we must either assume the solution takes the form of a plane wave or solve (14) numerically because the equation is nonseparable and the coefficients are nonconstant. We discuss our numerical solution to the problem in the next section. Here we develop an analytical expression for the wave-period-averaged energy density in terms of Ψ and then adopt the plane wave assumption, investigate the properties of the phase and group velocity, obtain a full set of polarization relations, and use these relations to derive the classical expression for the energy flux
a. Energy density





b. Plane wave dispersion and polarization relations
We now make the Wentzel–Kramers–Brillouin (WKB) approximation of geometrical optics (i.e., the plane wave assumption). Strictly speaking, this approximation is only valid when the wavelength is much smaller than the length scale of variations in the background mean flow (i.e., spatial variation in the wavelength occurs over much larger scales than the wavelength) (Bender and Orszag 1978). This is not necessarily true for the flows of interest in this problem. Moreover, ray tracing cannot accurately describe tunneling problems, where there is a small barrier region in which ω < ωmin, or scattering problems. Furthermore, the ray tracing solution has grave errors at turning points (where the governing equation switches from hyperbolic to elliptic behavior). Thus, we check the results with a numerical solution, the accuracy of which does not depend on this scale separation.
First, we assume that the solutions to (8)–(12) are all of the plane wave form. For example,

c. Phase and group velocity
We also note that the group velocity goes to zero where ω = ωmin, as shown, for example, in Fig. 11. Consequently, for two-dimensional waves, we expect that most of the wave energy will be trapped inside the region where ω > ωmin.
(top) Rays (parallel to characteristics) in the idealized background flow for waves with ω = 0.95f. The variation of the magnitude of the group velocity (normalized by its maximum along each ray) is colored on the rays. (middle) A numerical solution to Eq. (51) with ω = 0.95f and a point source at y ≈ 0.75 km, z ≈ −100 m. This is a snapshot of the streamfunction in time and the dashed cyan lines are the same rays shown in the top panel. (bottom) Energy density
Citation: Journal of Physical Oceanography 43, 4; 10.1175/JPO-D-12-0132.1
d. Energy flux


5. Ray tracing and numerical solution
We now discuss two ways to develop a solution to the energy propagation and amplification problem for NIWs governed by (14). In both cases we solve the wave problem in our idealized test environment described in section 3 and plotted in Figs. 2, 3, and 5. We begin by describing and presenting results from ray tracing. Then we compare the ray tracing results with a fourth-order accurate finite difference solution to (14).
a. Ray tracing
In this section, we describe the ray-tracing solution to (14) where we use the WKB approximation and assume that the solution can be described by a plane wave as described in section 4. Since the medium varies, this is an approximation everywhere in the domain. Nevertheless, it allows for a great deal of theoretical progress and, as we will see, this solution qualitatively reproduces the features of the numerical solution. Therefore, a thorough understanding of the properties of this solution aids the understanding of NIWs in baroclinic currents.
The results of some ray-tracing computations are shown in Figs. 5 and 11. Figure 5 shows some rays, parallel to characteristics (23), computed with baroclinicity (i.e., with the S2 term included) and without baroclinicity (i.e., with the S2 term set to zero) so as to highlight the qualitative modification of the ray structure due to baroclinicity. Figure 11, on the other hand, shows that the ray-tracing results with baroclinicity compare favorably with the numerical solution described below. In the top panel of Fig. 11, the colored scatter lines are again rays, parallel to characteristics. The color indicates the magnitude of the group velocity |cg|, normalized by the maximum of |cg| along each ray (the group velocity magnitude is symmetric about y = 0). The background flow shown in both Figs. 5 and 11 is the same idealized flow described throughout the paper and the wave packets are monochromatic, with a frequency ω = 0.95f. The reflection rule, applied at vertical and horizontal boundaries and turning points requires that the characteristic and direction of energy propagation switch.
We make several observations. First, the rays are constrained to lie in the region where ω ≥ ωmin. The group velocity (49) is not real outside this region and the governing equation is no longer hyperbolic as we have discussed. We also note that, as the ray approaches the magenta contour [where ω = ωmin and (14) is parabolic], the slope of the ray approaches the local slope of isopycnals, θb, as we would expect.
To further characterize the behavior of the solution near ωmin we use the terminology of “turning points” and “critical layers.” When we refer to behavior typical of a turning point, we mean that the behavior is qualitatively similar to the behavior at a turning point in a 1D Sturm–Liouville problem, defined rigorously in, for example, Bender and Orszag (1978). This is an internal reflection point in a ray theory where the group velocity goes to zero but multiple characteristics do not converge. More precisely, a turning point occurs where the slope of the ωmin contour is not parallel to characteristics. In contrast, a critical layer is defined as a region where many characteristic curves converge to the same point or line at ωmin. In this case, the characteristics are not only approaching the ωmin curve but their slopes are converging to the slope of this curve as well.
In general, the ωmin contour can act as both a turning point and a critical layer in the same background flow. To see this, consider our idealized domain. Except at two points in the legs of the magenta bowl (in Fig. 11), which bounds the region where ω > ωmin, the slope of this contour is not equal to the slope of characteristics or, equivalently, isopycnals. Therefore, away from these two points the ωmin contour is best characterized as a turning point because multiple characteristics do not converge. On the other hand, near these two points this contour is best characterized as a critical layer because all characteristics converge to one point here and the group velocity (parallel to characteristics) goes to zero.
Although the critical layer is localized to a point in the idealized flow (e.g., Fig. 5a), there is no particular requirement that a critical layer be a point. It merely requires that the ωmin boundary be locally parallel to characteristics or, equivalently, isopycnals. In the Gulf Stream flow, for example (in Fig. 5c), there are multiple critical layers where the ωmin contour is parallel to isopycnals. These regions coincide with the observed high ageostrophic shears, as shown in Fig. 1. We refer to these layers as slantwise critical layers because the isopycnals on which they are found are slanted. This phenomenon is analogous to the vertical critical layer described by Kunze (1985) where the relative vorticity of a mean flow increases with depth but buoyancy remains constant (see his Fig. 14). We adopt this new terminology to additionally distinguish slantwise critical layers from vertical critical layers that can form when inertia–gravity waves propagate parallel to a vertically sheared background flow (e.g., Booker and Bretherton 1967), in contrast to the case of normal incidence we consider here. The crucial difference between these two phenomena is that in a slantwise critical layer, as described here, the component of the wave’s phase velocity in the direction of the mean current is zero, so it can never equal the local velocity of the background flow. This means that a normally incident two-dimensional NIW in a steady two-dimensional background flow cannot transfer energy to the background flow by this process.
We use the fact that
b. Numerical solution



We discretize the domain with 802 points evenly distributed in the horizontal and vertical directions. We present this numerical solution to (51) in Fig. 11 for ω = 0.95f and a point source at approximately 100-m depth in the middle of the domain. We plot the real part of the streamfunction Ψ in the middle panel and the wave-period-averaged energy density
The computation (51) is not intensive; it takes only several seconds on a single core. Experiments with various levels of friction as well as nontraditional and nonhydrostatic effects were performed but are not presented. A certain amount of friction is required to damp very small-scale features. Without any friction, small-scale features with unrealistically large shears can exist at or near the grid scale. However, adding nonhydrostatic and nontraditional terms does not qualitatively change the results.
The results of the numerical computation show striking similarities to the ray tracing despite the fact that the medium varies relatively rapidly in space. Streamlines, high gradients in Ψ, and high energy densities occur along rays. Streamlines are nearly parallel to isopycnals near ωmin (the magenta contour) and, as expected, the numerical solution is amplified there and decays outside, similar to an Airy function at a turning point. Furthermore, the strongest amplification occurs near the critical layers (the corners of the bowl where ωmin is parallel to isopycnals in Fig. 11). The shear, which is by definition parallel to rays throughout the domain, and, in particular, isopycnals near ωmin, is strongest near critical layers where wave velocities, energy densities, gradients in energy density, and hence shears are largest.
In short, the results of the ray tracing and numerical solution are consistent. Both are useful tools for analyzing waves in baroclinic currents. Regardless of which method is used, however, it is important to include the effects of baroclinicity when the Richardson number of the background flow is less than ~10.
6. Discussion




(a) The expansion parameter,
Citation: Journal of Physical Oceanography 43, 4; 10.1175/JPO-D-12-0132.1
As mentioned above, the theory presented originally by Mooers (1975) and more recently by Plougonven and Zeitlin (2005) is also appropriate in the strongly baroclinic regime. Our work is distinguished from that of both Mooers (1975) and Plougonven and Zeitlin (2005) in that we have emphasized the physical interpretation of NIWs rather than the mathematical properties of solutions to (14). We have also presented a numerical solution in an idealized jet with properties similar to the Gulf Stream to illustrate the results, highlight their applicability to real world flows and confirm the inferences derived from ray tracing and parcel arguments. We must also point out that the results of the hydrostatic theory presented in this paper and in Plougonven and Zeitlin are not entirely the same as those in Mooers. One particularly prominent difference is that the group velocity does not go to zero at ωmin if the nonhydrostatic equations are used (Mooers 1975). Nevertheless, this modification will have a negligible effect on the energy flux of NIWs in an environment like our model environment or the Gulf Stream. The group velocity still gets very small at this point and the hydrostatic approximation is appropriate for these low-aspect-ratio waves.
Plougonven and Zeitlin (2005) recently presented a similar hydrostatic formulation of the Eliassen–Sawyer equation. They used the theory to study geostrophic adjustment rather than wind-generated near-inertial waves, although they did touch upon the trapping of near-inertial waves in geostrophic jets. However, they analyzed the trapping phenomena using a barotropic form of (14); that is, they only considered variations in mean flow relative vorticity. Therefore, like Kunze (1985), they did not explore the baroclinic effects that we have shown are important.
Finally, we observe that the theory presented here is similar to the theory of NIWs on the nontraditional β plane (e.g., Gerkema and Shrira 2005b; Colin de Verdiere 2012). The horizontal variation in f due to changes in latitude on the nontraditional β plane is analogous to the horizontal variation in F due to variations in relative vorticity presented in this paper. As a result, on a β plane, NIWs have a turning latitude. However, because the horizontal component of the Coriolis frequency
7. Conclusions
In this paper, we have analyzed two-dimensional near-inertial waves (NIWs) that propagate perpendicular to baroclinic geostrophic currents with O(1) Rossby and Richardson numbers. The work builds primarily off that of Mooers (1975), who derived the mathematical properties of nonhydrostatic inertia–gravity waves in this system, but emphasizes a new physical interpretation for the results using parcel arguments and conservation principles. In particular, conservation of absolute momentum MT = u − fy is crucial for understanding the unusual properties of the waves in this system. One such unusual property of these waves is that their minimum frequency decreases with increasing baroclinicity. This can be understood geometrically. The waves of lowest frequency have parcel displacements along isopycnals and thus experience no buoyancy force. The restoring force for the oscillations in this case is provided solely by the Coriolis force, whose strength, by conservation of MT, depends on the along-isopycnal gradient of geostrophic absolute momentum, Mg = ug − fy. This gradient is reduced as baroclinicity increased because isopycnals steepen while Mg surfaces flatten, thereby lowering the minimum frequency.
By lowering the intrinsic frequency, baroclinicity allows NIWs to exist in regions where the effective Coriolis frequency,
Near-inertial waves can be trapped and amplified in the anomalously low frequency regions. The amplification preferentially occurs near slantwise critical layers that run parallel to isopycnals. These are places where multiple characteristics converge and the magnitude of the group velocity decays to zero (see e.g., Fig. 11). At these critical layers, the waves attain their minimum frequency and hence parcel displacements, rays, characteristics, and lines of constant phase all run parallel to isopycnals. Thus, near a slantwise critical layer, the vertical shear of the NIW would be strongest and lines of constant shear would tend to align with isopycnals. This characterization is consistent with the enhanced ageostrophic shear in the Gulf Stream highlighted in Fig. 1, which occurs near a region with multiple critical layers, as identified by a ray-tracing calculation that accounts for the effects of baroclinicity (see e.g., Fig. 5).
Another property of the anomalously low frequency waves that differs from classical inertia–gravity waves is their polarization relation. In particular, the waves with the lowest frequency are characterized by horizontal velocities that are not circularly polarized, unlike classical NIWs. As baroclinicity is increased, the polarization relation of these waves becomes more rectilinear, with stronger velocities in the cross- versus alongfront direction. As shown by Thomas (2012) this change in polarization relation can lead to an efficient exchange of kinetic energy between the waves and balanced flows in regions of active frontogenesis. The idea being that, as the polarization relation shifts from circular to rectilinear, the waves induce a momentum flux that is in a particular direction. When this momentum flux is pointed down the gradient in momentum of the balanced flow, the waves act as an effective viscosity, extracting kinetic energy from the mean flow.
The polarization relation is modified in unexpected ways for waves of higher frequencies as well. In the presence of baroclinicity, the sense of rotation and the ellipticity of the hodograph traced out by the velocity vector over time for waves of the same frequency but different characteristics (i.e., different propagation directions) are not necessarily the same. For the case when one of the characteristics is steeper than surfaces of constant Mg, the velocity vector for a wave in the Northern (Southern) Hemisphere rotates counterclockwise (clockwise) with time, that is, opposite to what classical theory would predict. This implies that the resonance conditions for maximal wind work on the near-inertial motions no longer corresponds to an anticyclonic rotary wind oscillating at the inertial frequency, like the classical prediction of Pollard and Millard (1970). How much of an effect this has on the generation of NIWs in the proximity of the fronts, especially those associated with separated western boundary currents that underlie the midlatitude storm tracks, is an open question and one that will be the subject of future research on the generation process through which we hope to better understand the rates of wind generation, radiative decay, trapping, and dissipation of near-inertial energy in strongly baroclinic western boundary currents.
Another topic that we are pursuing is how three-dimensional NIWs (i.e., waves that have alongfront variability) in a strongly baroclinic current differ from the 2D waves described here. When the waves vary in the alongfront direction, a Doppler shift is present that can modify the conditions for wave trapping (e.g., Kunze 1985). Determining how 3D dynamics alters the key properties of NIWs, not just their trapping, is the subject of a follow-up study to this work.
Acknowledgments
We would like to thank four anonymous reviewers for their helpful suggestions. This work was supported by the Office of Naval Research Grant N00014-09-1-0202 and the National Science Foundation Grant OCE-0961714.
REFERENCES
Alford, M., 2003: Redistribution of energy available for ocean mixing by long-range propagation of internal waves. Nature, 423, 159–163.
Balmforth, N. J., S. G. Llewellyn Smith, and W. R. Young, 1998: Enhanced dispersion of near-inertial waves in an idealized geostrophic flow. J. Mar. Res., 56, 1–40.
Bender, C., and S. Orszag, 1978: Advanced Mathematical Methods for Scientists and Engineers. McGraw Hill, 607 pp.
Booker, J. R., and F. P. Bretherton, 1967: The critical layer for internal gravity waves in a shear flow. J. Fluid Mech., 27, 513–539.
Chuang, W., and D. Wang, 1981: Effects of density front on the generation and propagation of internal tides. J. Phys. Oceanogr., 11, 1357–1374.
Colin de Verdiere, A., 2012: The stability of short symmetric internal waves on sloping fronts: Beyond the traditional approximation. J. Phys. Oceanogr., 42, 459–475.
D’Asaro, E., C. Lee, L. Rainville, R. Harcourt, and L. Thomas, 2011: Enhanced turbulence and energy dissipation at ocean fronts. Science, 332, 318–322.
Eliassen, A., 1962: On the vertical circulation in frontal zones. Geofys. Publ., 24, 147–160.
Federiuk, J., and J. S. Allen, 1996: Model studies of near-inertial waves in flow over the Oregon continental shelf. J. Phys. Oceanogr., 26, 2053–2075.
Ferrari, R., and C. Wunsch, 2009: Ocean circulation kinetic energy: Reservoirs, sources and sinks. Annu. Rev. Fluid Mech., 41, 253–282.
Fomin, L. M., 1973: Inertial oscillations in a horizontally inhomogeneous current velocity field. Izv. Atmos. Ocean. Phys., 9, 75–83.
Gerkema, T., and V. I. Shrira, 2005a: Near-inertial waves in the ocean: Beyond the “traditional approximation.” J. Fluid Mech., 529, 195–219.
Gerkema, T., and V. I. Shrira, 2005b: Near-inertial waves on the “non-traditional” β plane. J. Geophys. Res., 110, C01003, doi:10.1029/2004JC002519.
Hoskins, B., 1974: The role of potential vorticity in symmetric stability and instability. Quart. J. Roy. Meteor. Soc., 100, 480–482.
Hua, B., D. Moore, and S. Le-Gentil, 1997: Inertial nonlinear equilibration of equatorial flows. J. Fluid Mech., 331, 345–371.
Inoue, R., M. C. Gregg, and R. R. Harcourt, 2010: Mixing rates across the Gulf Stream, Part 1: On the formation of Eighteen Degree Water. J. Mar. Res., 68, 643–671.
Klein, P., and B. L. Hua, 1988: Mesoscale heterogeneity of the wind-driven mixed layer: Influence of a quasigeostrophic flow. J. Mar. Res., 46, 495–525.
Klein, P., and A. M. Treguier, 1995: Dispersion of wind induced inertial waves by a barotropic jet. J. Mar. Res., 53, 1–22.
Kunze, E., 1985: Near-inertial wave propagation in geostrophic shear. J. Phys. Oceanogr., 15, 544–565.
Kunze, E., and T. Sanford, 1984: Observations of near-inertial waves in a front. J. Phys. Oceanogr., 14, 566–581.
Kunze, E., R. W. Schmidt, and J. M. Toole, 1995: The energy balance in a warm core ring’s near-inertial critical layer. J. Phys. Oceanogr., 25, 942–957.
Leaman, K., and T. Sanford, 1975: Vertical energy propagation of inertial waves: A vector spectral analysis of velocity profiles. J. Geophys. Res., 80, 1975–1978.
Lee, D., and P. Niiler, 1998: The inertial chimney: The near-inertial energy drainage from the ocean surface to the deep layer. J. Geophys. Res., 103 (C4), 7579–7591.
Lighthill, J., 1978: Waves in Fluids. Cambridge University Press, 524 pp.
Magaard, L., 1968: Ein Beitrag zur Theorie der internen Wellen als Störungen geostrophischer Strömungen. Deut. Hydro. Zeit., 21, 241–278.
Marshall, J., and Coauthors, 2009: The CLIMODE field campaign: Observing the cycle of convection and restratification over the Gulf Stream. Bull. Amer. Meteor. Soc., 90, 1337–1350.
Mooers, C. N. K., 1970: The interaction of an internal tide with the frontal zone in a coastal upwelling region. Ph.D. thesis, Oregon State University, 480 pp.
Mooers, C. N. K., 1975: Several effects of a baroclinic current on the cross-stream propagation of inertial-internal waves. Geophys. Fluid Dyn., 6, 245–275.
Munk, W., and C. Wunsch, 1998: Abyssal recipes II: Energetics of tidal and wind mixing. Deep-Sea Res. I, 45, 1977–2010.
Nagai, T., A. Tandon, H. Yamazaki, and M. J. Doubell, 2009: Evidence of enhanced turbulent dissipation in the frontogenetic Kuroshio front thermocline. Geophys. Res. Lett., 36, L12609, doi: 10.1029/2009GL038832.
Perkins, H., 1976: Observed effect of an eddy on inertial oscillations. Deep-Sea Res., 23, 1037–1042.
Plougonven, R., and V. Zeitlin, 2005: Lagrangian approach to geostrophic adjustment of frontal anomalies in a stratified fluid. Geophys. Fluid Dyn., 99, 101–135.
Pollard, R. T., and R. C. Millard, 1970: Comparison between observed and simulated wind-generated inertial oscillations. Deep-Sea Res., 17, 813–821.
Rainville, L., and R. Pinkel, 2004: Observations of energetic high-wavenumber internal waves in the Kuroshio. J. Phys. Oceanogr., 34, 1495–1505.
Sawyer, J., 1956: The vertical circulation at meteorological fronts and its relation to frontogenesis. Proc. Roy. Soc. London, A234, 346–362.
Shcherbina, A. Y., L. D. Talley, E. Firing, and P. Hacker, 2003: Near-surface frontal zone trapping and deep upward propagation of internal wave energy in the Japan East Sea. J. Phys. Oceanogr., 33, 900–912.
Thomas, L., 2012: On the effects of frontogenetic strain on symmetric instability and inertiagravity waves. J. Fluid Mech.,711, 620–640, doi:10.1017/jfm.2012.416.
Thomas, L., and C. Lee, 2005: Intensification of ocean fronts by down-front winds. J. Phys. Oceanogr., 35, 1086–1102.
Thomas, L., and T. Joyce, 2009: Subduction on the northern and southern flanks of the Gulf Stream. J. Phys. Oceanogr., 40, 429–438.
Thomas, L., C. Lee, and Y. Yoshikawa, 2010: The subpolar front of the Japan/East Sea. Part II: Inverse method for determining the frontal vertical circulation. J. Phys. Oceanogr., 40, 3–25.
Thomas, L., J. Taylor, R. Ferrari, and T. Joyce, 2013: Symmetric instability in the Gulf Stream. Deep-Sea Res. II, in press.
Wang, D. P., 1991: Generation and propagation of inertial waves in the subtropical front. J. Mar. Res., 49, 619–633.
Weller, R. A., 1982: The relation of near-inertial motions observed in the mixed layer during the Jasin (1978) experiment to the local wind stress and to quasi-geostrophic flow field. J. Phys. Oceanogr., 12, 1122–1136.
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, doi:10.1029/2012GL053196.
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.
Wunsch, C., and R. Ferrari, 2004: Vertical mixing, energy, and the general circulation of the oceans. Annu. Rev. Fluid Mech., 36, 281–314.
Young, W. R., and M. Ben-Jelloul, 1997: Propagation of near-inertial oscillations through a geostrophic flow. J. Mar. Res., 55, 735–766.
Zeitlin, V., G. M. Reznik, and M. Ben Jelloul, 2003: Nonlinear theory of geostrophic adjustment. Part 2. Two layer and continuously stratified primitive equations. J. Fluid Mech., 491, 207–228.
Geostrophic flows with extremely high baroclinicity can develop enhanced shear and dissipation parallel to isopycnals even without significant inertial wave activity. Frontogenetic strain, in particular, can play an important role in this process and can also modify any internal waves in the front (Thomas 2012). However, frontogenesis does not appear to be important in the observations presented here and is not the focus of the present paper. See Winkel et al. (2002), Nagai et al. (2009), Thomas et al. (2010), and D’Asaro et al. (2011) for other examples of enhanced shear parallel to isopycnals due to waves and other phenomena.
The selected frequencies are somewhat arbitrary but, when ω is not too far from f and the flow is baroclinic, the observed separation between ωmin and F is typical. The qualitative nature of the results does not change for slightly different frequencies [i.e., (0.95 ± 0.05)f]. However, waves with much larger frequencies are not trapped whereas waves with much smaller frequencies cannot exist.
The theory of Young and Ben-Jelloul (1997) used in, for example, Balmforth et al. (1998) and Zeitlin et al. (2003) depends on an asymptotic expansion in small