1. Introduction
Inverse models of the ocean’s circulation suggest that much of the zonally integrated diapycnal upwelling that closes the abyssal meridional overturning circulation (AMOC) occurs in the tropical oceans (Lumpkin and Speer 2007). This result implies that intense mixing takes place in these regions. However, the abyssal near-equatorial oceans are poorly sampled; therefore, the inferences from the inverse models have not been fully tested in the field.
To address this problem and study mixing in the deep equatorial regions, Holmes et al. (2016, hereafter H16) recently carried out observations in the eastern equatorial Pacific. During this field campaign, a lowered acoustic Doppler current profiler (LADCP), a conductivity–temperature–depth (CTD) cage, and microstructure sensors were used to gather information on flow velocity, stratification of the water column, and turbulence, respectively. The instruments were deployed from the surface all the way to the seafloor. Following this procedure, H16 reported evidence of strong turbulence around 0.5°S, ranging from the seafloor up to 700 m above. The rate of kinetic energy dissipation associated with this turbulence had values typical of flows over rough topography, such as midocean ridges (based on values reported by Waterhouse et al. 2014). In H16, however, the observations were carried out over smooth topography, in a region where the generation of internal tides through tidal–topographic interactions is relatively weak (St. Laurent and Garrett 2002). Hence, the common interpretation that bottom-intensified mixing is driven by internal waves breaking locally through interactions with rough topography is not applicable in the region studied by H16.
H16 also reported evidence of a downward-propagating Yanai wave in the LADCP data with a downward energy flux of about 2 mW m−2. This rate is comparable to the 1 mW m−2 average dissipation rate estimated in the abyss in the mixing region. Therefore, H16 suggested that the mixing may have been energized by an equatorially trapped wave (ETW) whose physics had been modified by the horizontal component of Earth’s rotation
However, it remains to be seen whether this mechanism can explain the enhanced mixing observed by H16 because their calculation used several assumptions that may not be valid. Namely, the solutions to the wave equation described by H16 assume that the waves are linear and two-dimensional. Nonlinear dynamics are known to be dominant at critical reflections (Dauxois and Young 1999). As a result, they might play a significant role around the inertial latitude when nontraditional effects are permitted. In particular, these processes could act to enhance or suppress the mixing that develops or to modify its spatial structure. In addition, the two-dimensional wave equation that H16 considered fails to describe the equatorial dynamics accurately. This equation, by assuming zonally invariant motions, captures the physics of traveling waves confined to the meridional plane. It is thus suitable to study IGWs at midlatitudes, which have a rather high frequency (close to the inertial frequency in their source region) and a long zonal wavelength relative to their meridional wavelength. These waves propagate mainly toward the equator, eventually breaking or dissipating along the way (Alford et al. 2016). In contrast, at the equator, most of the variability in the deep is supported by waves that have a rather large meridional scale and that primarily propagate zonally over a much longer time period. In this study, we refer to such waves as ETWs in contrast to the midlatitude IGWs because they have fundamentally distinct physical properties (Philander 1978).
The properties of ETWs were first described in detail by Matsuno (1966) and Blandford (1966), and they have subsequently been observed and amply documented (see Wunsch and Gill 1976; Harvey and Patzert 1976; Weisberg and Horigan 1981; Brandt and Eden 2005; Farrar 2008; Smyth et al. 2015). In particular, Eriksen (1980) and Eriksen and Richman (1988) used time series from moored arrays in the central Pacific at depths of 1500 and 3000 m to provide one of the most comprehensive descriptions of the deep equatorial ocean. They found that the energy at these depths is distributed over a broad band in frequency, with wave periods ranging from days to months, but a narrow band in zonal wavenumber of order O(1000) km. In a recent review of the characteristics of the deep equatorial variability, Ascani et al. (2010) showed that these observations were consistent with results from several ocean general circulation models. Nevertheless, the propensity of these ETWs to drive mixing in the abyss when
The scant attention in the literature to the nontraditional effects on ETWs in comparison to midlatitude IGWs is somewhat surprising. At the equator,
In the study presented here, we investigate the influence of
The remainder of this paper is organized as follows. In section 2, we summarize earlier theoretical contributions on nontraditional effects on ETWs and discuss the conceptual differences with the midlatitude IGW case. We then comment on the implications that nontraditional effects have for wave reflection off a flat bottom. In section 3, we introduce the numerical model and describe the design of the numerical experiments. In section 4, we report the results from the numerical experiments and compare runs with and without nontraditional effects, highlighting the spatiotemporal characteristics of the mixing that develop when nontraditional effects are active and its implication for diapycnal transport in the abyss. In section 5, we interpret the results using two-dimensional, viscous, analytical models to describe how critical reflection is responsible for the enhanced shear when nontraditional effects are active. Then, we discuss the dependence of this mechanism on the key parameters of the waves and of the medium in which they propagate, and we highlight some crucial differences across ETW types that might explain the differences in mixing that we observe in the numerical simulations. Section 6 concludes by discussing the strengths and limitations of the theory and potential topics for future research.
2. Analytical solution of ETWs accounting for nontraditional effects
We begin with a brief overview of linear equatorial wave theory when
Note that these turning latitudes are different from the ones derived in the 2D case. In the 2D traditional case, the turning latitude is the inertial latitude (i.e.,
In these calculations, we have shown that even if the dispersion relation of the wave is not affected by
In the ocean, we expect the tilting of the phase lines due to
3. Numerical experiments
a. Model equations
b. Model setup
In our simulations, we use a 3D Cartesian coordinate system centered at the equator, extending up to about ±8° in latitude and 4500 m deep with a flat bottom. Two sponge layers are used on the southern and northern boundaries, ranging from ±5.5° to ±8°, where the velocity is smoothly damped to zero and the stratification relaxed to the initial conditions. These regions act to absorb weak nonlinear perturbations only because the linear dynamics considered here are confined to the equatorial waveguide. Similar to the analytic derivation in section 2, periodic boundary conditions are used in the zonal direction. At the bottom, a free-slip boundary condition is used to avoid the interference of mixing induced by drag with mixing induced by nontraditional effects. However, as discussed in section 5, considering a nonzero drag coefficient at the bottom yields similar conclusions to those that we will present in the next section.
The resolution of the model is (
In the model, salinity is constant in the whole domain. As a result, temperature alone governs buoyancy through a linear equation of state, such that
c. Wave generation
To mimic the generation of a downward-propagating ETW beam by upper-ocean processes, we relax the velocity and density fields in the upper part of the domain to the analytical solution of nontraditional, downward-propagating ETWs; that is, the
In our study, the mask
d. Experimental design
We ran a suite of numerical simulations with five ETWs of a variety of frequencies: a mode-1 Rossby wave, a Kelvin wave, eastward- and westward-propagating Yanai waves, and a mode-1 westward-propagating equatorial inertia–gravity wave. Here, “mode” refers to the meridional mode number of the wave, which is represented by n in Eq. (18). Since these waves arise from different physical mechanisms (e.g., from buoyancy restoring forces to potential vorticity conservation), we can explore how the modifications induced by
For each type of wave, we chose the same zonal and vertical wavelengths, and we allowed the frequency to vary according to the dispersion relation of the wave [Eq. (17)]. We set the zonal wavelength to 1200 km and the vertical wavelength to 1700 m, both of which are representative of the deep equatorial ocean [see Ascani et al. (2010) for a thorough description of ETWs’ length scales]. For each simulation, the wave is forced from the beginning of the simulation for up to 2 years.
Table 1 summarizes the key properties of the simulated waves in the generation layer, where the stratification is constant. Note the gap between the inertial and turning latitudes of the wave, even in the generation layer where stratification is rather high. This discrepancy is due to the zonal wavenumber of the wave, highlighting the difference between 2D and 3D wave physics. Below the generation layer, the turning latitude increases due to lower stratification values, while the inertial latitude remains constant. Note also that the Rossby and westward-propagating Yanai waves have the largest kinetic energy density, which is approximately one order of magnitude larger than the other waves. These waves are thus more prone to nonlinearities, as shown by their large
Properties of the ETWs generated in the numerical experiments. These properties are valid only in the upper layer, where the wave is generated and where the stratification is constant. Below this layer, the variations in stratification modify the properties of the wave. The period and zonal wavelength will, however, remain constant. For the Yanai wave, both westward- (W) and eastward- (E) propagating waves are simulated. The nondimensional parameters
4. Numerical solutions
a. Beams at the inertial latitude
Figure 3 shows snapshots of the resulting meridional velocity from each simulation, with and without
The Kelvin wave case is particularly compelling because the meridional velocity is everywhere zero under the TA (except right at the surface, since our solution for a downward-propagating wave does not satisfy the free-surface boundary condition). As a result, the sharp features are particularly visible in the Kelvin wave simulation where they form two wave beams. Figure 4 shows the time development of the beams at four different stages in the simulation. Once the wave reaches the bottom, enhanced velocity values appear close to the seafloor around the inertial latitude of the wave. As the simulation moves forward, these features intensify and start to form two well-defined beams that emanate from the inertial latitude. The beams then propagate upward in the water column, reflecting back and forth at the inertial latitude of the wave. Eventually, they reach the surface generation layer, where they are damped by the nudging terms.
The beams that appear in the nontraditional runs propagate in the meridional plane, contrary to the ETWs that form a standing wave in the meridional plane. In addition, the alternating jets that develop in the abyss poleward of the inertial latitude (visible in Figs. 3, 4) indicate that the waves propagate past the inertial latitude, as predicted by the 2D nontraditional theory for IGWs (Gerkema and Shrira 2005). This suggests that the sharp beams are governed by 2D physics contrary to the ETWs that force the simulations. For 2D waves (i.e.,
Figure 5 shows how the characteristics in the nontraditional case deviate from the characteristics in the traditional case where
In the Rossby wave simulation in Fig. 3, beams are not clearly visible, but nonetheless features with small vertical and meridional scales develop in the deep. These features are confined to latitudes below the inertial latitude, which is quite close to the equator, given the low frequency of the wave.
b. Mean flows and higher harmonics
In our simulations, nontraditional effects significantly affect the spectral signature of the ETWs. This is illustrated in Fig. 6, which shows the spectrum for the westward-propagating Yanai wave in frequency–meridional wavenumber space, averaged over the last 1500 m of the water column, below the generation layer. In the traditional run, the spectral energy is confined to the frequency and meridional scale of the simulated wave. In the nontraditional run, however, the spectral signature is significantly broadened to higher meridional wavenumbers, reflecting the presence of the beams. In addition, higher harmonics in frequency are evident and could play a significant role in the propagation of wave energy to higher latitudes since these waves of higher frequency are not necessarily confined to the inertial latitude of the forced wave. Note, however, that despite the energy in the higher harmonics, internal wave beams at those frequencies are not clearly visible in the solutions shown in Fig. 3. It is probable that the spatial structure of the higher harmonics does not satisfy the dispersion relation for inertia–gravity waves and therefore does not generate freely propagating waves.
It is also interesting to notice the enhanced power at
c. Abyssal mixing
The small-scale features that develop when
In all the simulations, the near-bottom enhancement of the diffusivity is zonally invariant (Fig. 9a) and sets up once the ETW reaches the bottom, having propagated down from the upper ocean. It persists as long as the forcing of the ETW is maintained, suggesting that no negative feedbacks act to weaken the mixing despite the enhancement of nonlinear effects due to
Figure 8 shows that the magnitude of the diffusivity and its spatial variability differ for each wave type. In particular, the westward inertia–gravity and eastward Yanai waves yield mixing regions with a large meridional extent. The Rossby wave, however, yields a mixing region with a higher vertical extent. In contrast, the westward Yanai wave induces lower diffusivities, suggesting that this wave type is less efficient at driving mixing.
d. Implications for diapycnal transport in the abyss
Figure 10 shows Hovmöller diagrams of
5. Beam formation mechanism
In the numerical solutions described in the previous section, the mixing in the abyss develops around the inertial latitude, where the wave beams originate, highlighting how the beams are essential for enhancing the mixing through strengthening the shear and lowering the Richardson number of the flow. In this section, we explore the formation mechanism of the beams to understand the key parameters that set their strength. We start by showing that the beams are excited because of the nonzero vertical velocity of the downward-propagating ETWs at the seafloor when nontraditional effects are active. Then, we focus on the region around the inertial latitude of the wave to show that a vertical velocity there can drive critical reflection of the ETW off the flat seafloor, which is permitted when
a. Beams generation through bottom reflection of ETWs
In section 2, we demonstrated that
To illustrate that this process can generate the beams that we observe in the numerical simulations, we solved the 2D wave [Eq. (12)] subject to a nonzero vertical velocity at the bottom. More specifically, we set the vertical velocity at the bottom equal to
Figure 11 shows the result for the case where we set
This calculation confirms that it is the change in the bottom reflection process induced by nontraditional effects that gives rise to the beams and that it does not depend on the specific details of the ETW’s spatial structure. Our interpretation is that the reflection of an ETW off a flat-bottom boundary in the nontraditional case requires the generation of higher meridional modes that are ultimately governed by the 2D physics and thus amplified at the inertial latitude.
b. Critical reflection around the inertial latitude
Figure 12 shows the characteristic curves in the three limits where
We conclude that the beams form through this process since they are strongest near the inertial latitude and since they run parallel to the shallow characteristic
c. Analytical solution for the reflected wave accounting for viscosity
This equation admits six solutions with roots
As illustrated in Fig. 14,
The KE and streamlines of the incident, reflected, and total solutions are shown in Fig. 15 for no-slip and free-slip bottom boundary conditions. The wave field consists of the incident wave with vertical wavenumber
In Fig. 16, we show the sensitivity of this ratio to different parameters: the latitude (through varying f ), the stratification, the viscosity, the period of the incident wave, and its vertical wavelength
d. Impact of the local energy flux
The KE of the reflected wave and hence the beams at the inertial latitude is a function of both the amplification factor in Eq. (41) and the KE of the incident wave KEinc. The latter is set by the local energy flux at the inertial latitude coming from the downward-propagating ETW. The various ETW types have different meridional structures and, consequently, different local energy fluxes even if they have the same meridionally averaged downward flux of energy. Consequently, the variations in the strength of the beams and mixing with wave type seen in Figs. 3 and 8 could arise from the sensitivity of the amplification factor to both wave properties, as discussed above, and differences in the local energy flux for each wave type.
To illustrate this idea, we use the local energy flux evaluated at the inertial latitude in the abyss from the simulations of ETWs shown in Fig. 3, and we derive the corresponding KE of the incident wave using the vertical group velocity of the ETWs [i.e.,
6. Discussion and conclusions
The current paradigm for the diapycnal upwelling of the AMOC attributes the mixing necessary to drive diapycnal motions to internal waves breaking over rough topography (Armi 1978; Polzin et al. 1997; Waterhouse et al. 2014; Ferrari 2014). Recent studies have emphasized the role of mixing on sloping topography, such as the boundaries of ocean basins, in facilitating diapycnal upwelling in the deep (McDougall and Ferrari 2017; Callies and Ferrari 2018). Here, we illustrate the potential importance of ETWs in driving diapycnal upwelling in the abyssal equatorial oceans, a feature that is seen in inverse calculations of the global meridional overturning circulation (Lumpkin and Speer 2007), and we explain the mechanism by which ETWs generate near-bottom mixing.
Using numerical simulations and analytical theory, we have shown that nontraditional effects induce a fundamental change in the bottom reflection of the ETWs. In the weakly stratified abyssal equatorial ocean, the meridional and vertical structures of ETWs are shaped by absolute momentum surfaces, which are curved when
While the diffusivities are bottom intensified, they induce diapycnal upwelling, not downwelling, because the stratification increases with height off the bottom, resulting in a convergent buoyancy flux. Integrated over the tropics and zonally over a typical wavelength of an ETW, we estimate that the mixing could yield of order 10 Sv of diapycnal transport in the few hundred meters above the seafloor and hence could contribute significantly to the diapycnal upwelling of the abyssal equatorial circulation inferred by inverse models. Our study is, however, potentially limited by the Richardson number–dependent mixing scheme that we used, which parameterizes the strength of the turbulent diffusivity. Having said that, the maximum value of the diffusivity employed in the mixing scheme is similar to estimates made from microstructure observations in the abyssal equatorial ocean (e.g., H16), suggesting that diffusivities in our simulations are not unrealistically large. Because the shape of the stratification profile in the abyss is crucial for the resulting upwelling, it is also necessary to investigate whether the feature shown in Fig. 2 is ubiquitous in the equatorial oceans.
The diffusivities that we see in our simulations are consistent with the localized mixing observed by H16 at 0.5°S, which corresponds to the inertial latitude of the Rossby wave in our simulations. However, H16 observed a mixing region ranging up to 700 m above the seafloor, while mixing patterns in our simulations extend up to 300 m above the seafloor. This discrepancy could be because H16 observed the mixing associated with a particular wave event, whereas we considered ETWs with energy fluxes consistent with the mean state of the equatorial ocean in our calculations.
In addition, the beams are also responsible for the generation of mean flows. This mean flow generation mechanism could influence the complex equatorial deep jets’ structure, whose origin is still debated (Luyten and Swallow 1976; Ménesguen et al. 2009; Ascani et al. 2015; Cravatte et al. 2017). Further work will investigate the link between nontraditional effects and the steady equatorial circulation.
In our simulations, we use an ETW wave maker that is zonally and temporally invariant and symmetric relative to the equator. In the real ocean, however, there is a large meridional asymmetry in the ETW field that can result in “shadow zones” where the wave energy does not converge (Schopf et al. 1981; Cane and Moore 1981). Similarly, zonal variations in the processes that generate ETWs, such as tropical instability waves (Eriksen and Richman 1988) or winds (Durland and Farrar 2012), can restrict the longitudinal extent of the ETW energy reaching the bottom (e.g., Ascani et al. 2010). Furthermore, these processes can vary on seasonal time scales and longer (e.g., associated with ENSO cycles). How these meridional, zonal, and temporal variations in the ETW wave field modulate the near-bottom mixing and the net diapycnal transport across the entire extent of the equatorial Pacific remains to be quantified. Future work will focus on running realistic simulations of the eastern equatorial Pacific with
Acknowledgments
This work was funded by the National Science Foundation Grant OCE-1260312. Thanks to T. Durland for his insightful comments and encouragements that have significantly improved the quality of this article. This work benefitted also from stimulating discussions with D. Moore, M. Fruman, J. Wenegrat, and L. Maas.
REFERENCES
Adcroft, A., and Coauthors, 2016: MITgcm user manual. MIT Tech. Rep., 451 pp., http://mitgcm.org/public/r2_manual/final/online_documents/manual.html.
Alford, M. H., J. A. MacKinnon, H. L. Simmons, and J. D. Nash, 2016: Near-inertial internal gravity waves in the ocean. Annu. Rev. Mar. Sci., 8, 95–123, https://doi.org/10.1146/annurev-marine-010814-015746.
Armi, L., 1978: Some evidence for boundary mixing in the deep ocean. J. Geophys. Res., 83, 1971–1979, https://doi.org/10.1029/JC083iC04p01971.
Ascani, F., E. Firing, P. Dutrieux, J. P. McCreary, and A. Ishida, 2010: Deep equatorial ocean circulation induced by a forced–dissipated Yanai beam. J. Phys. Oceanogr., 40, 1118–1142, https://doi.org/10.1175/2010JPO4356.1.
Ascani, F., E. Firing, J. P. McCreary, P. Brandt, and R. J. Greatbatch, 2015: The deep equatorial ocean circulation in wind-forced numerical solutions. J. Phys. Oceanogr., 45, 1709–1734, https://doi.org/10.1175/JPO-D-14-0171.1.
Blandford, R., 1966: Mixed gravity-Rossby waves in the ocean. Deep-Sea Res. Oceanogr. Abstr., 13, 941–961, https://doi.org/10.1016/0011-7471(76)90912-8.
Boyer, T. P., and Coauthors, 2013: World Ocean Database 2013. NOAA Tech. Rep. NESDIS 72, 208 pp., https://doi.org/10.7289/V5NZ85MT.
Brandt, P., and C. Eden, 2005: Annual cycle and interannual variability of the mid-depth tropical Atlantic Ocean. Deep-Sea Res. I, 52, 199–219, https://doi.org/10.1016/j.dsr.2004.03.011.
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.
Cane, M. A., and D. W. Moore, 1981: A note on low-frequency equatorial basin modes. J. Phys. Oceanogr., 11, 1578–1584, https://doi.org/10.1175/1520-0485(1981)011<1578:ANOLFE>2.0.CO;2.
Cox, M., 1980: Generation and propagation of 30-day waves in a numerical model of the Pacific. J. Phys. Oceanogr., 10, 1168–1186, https://doi.org/10.1175/1520-0485(1980)010<1168:GAPODW>2.0.CO;2.
Cravatte, S., E. Kestenare, F. Marin, P. Dutrieux, and E. Firing, 2017: Subthermocline and intermediate zonal currents in the tropical Pacific Ocean: Paths and vertical structure. J. Phys. Oceanogr., 47, 2305–2324, https://doi.org/10.1175/JPO-D-17-0043.1.
Dauxois, T., and W. R. Young, 1999: Near-critical reflection of internal waves. J. Fluid Mech., 390, 271–295, https://doi.org/10.1017/S0022112099005108.
Durland, T. S., and J. T. Farrar, 2012: The wavenumber–frequency content of resonantly excited equatorial waves. J. Phys. Oceanogr., 42, 1834–1858, https://doi.org/10.1175/JPO-D-11-0234.1.
Eriksen, C. C., 1980: Evidence for a continuous spectrum of equatorial waves in the Indian Ocean. J. Geophys. Res., 85, 3285–3303, https://doi.org/10.1029/JC085iC06p03285.
Eriksen, C. C., 1981: Deep currents and their interpretation as equatorial waves in the western Pacific Ocean. J. Phys. Oceanogr., 11, 48–70, https://doi.org/10.1175/1520-0485(1981)011<0048:DCATIA>2.0.CO;2.
Eriksen, C. C., and J. G. Richman, 1988: An estimate of equatorial wave energy flux at 9- to 90-day periods in the central Pacific. J. Geophys. Res., 93, 15 455–15 466, https://doi.org/10.1029/JC093iC12p15455.
Farrar, J. T., 2008: Observations of the dispersion characteristics and meridional sea level structure of equatorial waves in the Pacific Ocean. J. Phys. Oceanogr., 38, 1669–1689, https://doi.org/10.1175/2007JPO3890.1.
Feistel, R., 2008: A Gibbs function for seawater thermodynamics for −6 to 80°C and salinity up to 120 g kg−1. Deep-Sea Res. I, 55, 1639–1671, https://doi.org/10.1016/j.dsr.2008.07.004.
Ferrari, R., 2014: What goes down must come up. Nature, 513, 179–180, https://doi.org/10.1038/513179a.
Fruman, M. D., 2009: Equatorially bounded zonally propagating linear waves on a generalized β plane. J. Atmos. Sci., 66, 2937–2945, https://doi.org/10.1175/2009JAS2932.1.
Gerkema, T., and V. I. Shrira, 2005: Near-inertial waves on the “nontraditional” β plane. J. Geophys. Res., 110, C01003, https://doi.org/10.1029/2004JC002519.
Harvey, R., and W. Patzert, 1976: Deep current measurements suggest long waves in the eastern equatorial Pacific. Science, 193, 883–885, https://doi.org/10.1126/science.193.4256.883.
Holmes, R. M., J. N. Moum, and L. N. Thomas, 2016: Evidence for seafloor-intensified mixing by surface-generated equatorial waves. Geophys. Res. Lett., 43, 1202–1210, https://doi.org/10.1002/2015GL066472.
Knox, R., and D. L. Anderson, 1985: Recent advances in the study of the low-latitude ocean circulation. Prog. Oceanogr., 14, 259–317, https://doi.org/10.1016/0079-6611(85)90014-X.
Large, W. G., J. C. McWilliams, and S. C. Doney, 1994: Oceanic vertical mixing: A review and a model with a nonlocal boundary layer parameterization. Rev. Geophys., 32, 363–403, https://doi.org/10.1029/94RG01872.
Lumpkin, R., and K. Speer, 2007: Global ocean meridional overturning. J. Phys. Oceanogr., 37, 2550–2562, https://doi.org/10.1175/JPO3130.1.
Luyten, J. R., and J. C. Swallow, 1976: Equatorial undercurrents. Deep-Sea Res. Oceanogr. Abstr., 23, 999–1001, https://doi.org/10.1016/0011-7471(76)90830-5.
Marshall, J., C. Hill, L. Perelman, and A. Adcroft, 1997: Hydrostatic, quasi-hydrostatic, and nonhydrostatic ocean modeling. J. Geophys. Res., 102, 5733–5752, https://doi.org/10.1029/96JC02776.
Matsuno, T., 1966: Quasi-geostrophic motions in the equatorial area. J. Meteor. Soc. Japan, 44, 25–43, https://doi.org/10.2151/jmsj1965.44.1_25.
McCreary, J. P., 1984: Equatorial beams. J. Mar. Res., 42, 395–430, https://doi.org/10.1357/002224084788502792.
McDougall, T. J., 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.
Ménesguen, C., B. L. Hua, M. D. Fruman, and R. Schopp, 2009: Dynamics of the combined extra-equatorial and equatorial deep jets in the Atlantic. J. Mar. Res., 67, 323–346, https://doi.org/10.1357/002224009789954766.
Needler, G., and P. Leblond, 1973: On the influence of the horizontal component of the Earth’s rotation on long period waves. Geophys. Fluid Dyn., 5, 23–45, https://doi.org/10.1080/03091927308236107.
Philander, S. G., 1978: Forced oceanic waves. Rev. Geophys., 16, 15–46, https://doi.org/10.1029/RG016i001p00015.
Polzin, K., J. 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.
Raymond, W. H., 2000: Equatorial meridional flows: Rotationally induced circulations. Pure Appl. Geophys., 157, 1767–1779, https://doi.org/10.1007/PL00001060.
Raymond, W. H., 2001: Kelvin waves: Rotationally induced circulations. Dyn. Atmos. Oceans, 34, 23–43, https://doi.org/10.1016/S0377-0265(01)00058-6.
Roundy, P. E., and M. A. Janiga, 2012: Analysis of vertically propagating convectively coupled equatorial waves using observations and a non-hydrostatic Boussinesq model on the equatorial beta-plane. Quart. J. Roy. Meteor. Soc., 138, 1004–1017, https://doi.org/10.1002/qj.983.
Schopf, P. S., D. L. Anderson, and R. Smith, 1981: Beta-dispersion of low-frequency Rossby waves. Dyn. Atmos. Oceans, 5, 187–214, https://doi.org/10.1016/0377-0265(81)90011-7.
Smyth, W. D., T. S. Durland, and J. N. Moum, 2015: Energy and heat fluxes due to vertically propagating Yanai waves observed in the equatorial Indian Ocean. J. Geophys. Res. Oceans, 120, 6630–6647, https://doi.org/10.1002/2014JC010152.
St. Laurent, L., and C. Garrett, 2002: The role of internal tides in mixing the deep ocean. J. Phys. Oceanogr., 32, 2882–2899, https://doi.org/10.1175/1520-0485(2002)032<2882:TROITI>2.0.CO;2.
Waterhouse, A. F., and Coauthors, 2014: Global patterns of diapycnal mixing from measurements of the turbulent dissipation rate. J. Phys. Oceanogr., 44, 1854–1872, https://doi.org/10.1175/JPO-D-13-0104.1.
Weisberg, R. H., and A. M. Horigan, 1981: Low-frequency variability in the equatorial Atlantic. J. Phys. Oceanogr., 11, 913–920, https://doi.org/10.1175/1520-0485(1981)011<0913:LFVITE>2.0.CO;2.
Winters, K. B., P. Bouruet-Aubertot, and T. Gerkema, 2011: Critical reflection and abyssal trapping of near-inertial waves on a β-plane. J. Fluid Mech., 684, 111–136, https://doi.org/10.1017/jfm.2011.280.
Wunsch, C., and A. E. Gill, 1976: Observations of equatorially trapped waves in Pacific sea level variations. Deep-Sea Res. Oceanogr. Abstr., 23, 371–390, https://doi.org/10.1016/0011-7471(76)90835-4.