1. Introduction
The initial response of the ocean mixed layer to any change in the wind stress may be approximated using a one-dimensional slab model (Pollard and Millard 1970; Pollard 1980) in which momentum is uniform within the mixed layer. The momentum balance may be split into an Ekman contribution, with the wind stress balanced by part of the Coriolis force, and an inertial oscillation (IO), in which local acceleration balances the remainder of the Coriolis force. In the IO, the velocity vector rotates anticyclonically at the frequency f (the Coriolis parameter). The slab model may be improved by taking into account a finite domain, horizontal variations in the wind, and the variation of f with latitude, any of which introduces horizontal divergence and horizontal pressure gradients (D’Asaro 1989). The inertial momentum balance is modified by the addition of the horizontal pressure gradient to yield a rotationally modified gravity wave, or Poincaré wave, with frequency f as the lower limit. Hence, the rotation frequency of the velocity vector shifts to slightly higher frequencies (i.e., it is blue shifted), and the motion can be described as a near-inertial oscillation (NIO). When the propagation of these waves is important, we may alternatively call them near-inertial waves (NIWs). In a model with continuous stratification below the mixed layer, NIW propagation is fully three-dimensional.
Kundu et al. (1983) explored the behavior of wind-generated currents along an infinite, vertical wall, with continuous stratification below a surface mixed layer. Their solution to the linear problem was expressed as the sum of Ekman and NIO responses in the mixed layer, plus NIW energy propagating both offshore and downward. The net effect is a solution for which the amplitude of the mixed layer NIO tends to zero at the coast, consistent with the no-normal-flow boundary condition. This phenomenon, known as coastal inhibition, has been well documented and observed (Schahinger 1988; Tintoré et al. 1995; Shearman 2005; Fontán and Cornuelle 2015).
The occurrence of coastal inhibition near islands is unclear, as islands have both finite lateral scale and complex bathymetric profiles. Few studies have explored the interaction of NIOs with islands, though in a recent paper, Brink (2021) modeled the island response to large-scale currents oscillating at tidal frequencies. He found that superinertial “leaky” modes can be generated; they resemble Kelvin waves propagating around the island, but they slowly lose energy to internal waves that radiate away.
In the second paper, Longuet-Higgins (1970, hereafter LH70) focused on inertial motions around circular and elliptical islands. By setting σ = −f and applying the boundary condition that the velocity tends to an IO at infinity, the long-wave equation (1) is reduced to the Laplace equation. In contrast to the coastal inhibition observed along an infinite coastal boundary, Longuet-Higgins found that inertial currents strengthen near the boundary of a circular island, reaching speeds twice that of the far-field IOs. With an elliptical island, inertial currents are further enhanced near the narrow tips of the island. While LH69 considered the trapped wave modes of an island, the solutions of LH70 can be thought of as the adjustment to an IO near an island boundary and are relevant at the inertial frequency around an island with a radius much smaller than the Rossby radius of deformation.
In an observational study, Siegelman et al. (2019) examined near-inertial currents around the island chain of Palau. Near-inertial motions coherent between the northern and southern tips of the island chain were observed, despite over 170 km separating the two sites. The generation of these coherent motions was attributed to wind stress, which had decorrelation length scales of 380 km. Spatial variability of the near-inertial currents around the island chain also was observed, with stronger currents near the narrow northern and southern tips of the island than in the far field, which is qualitatively consistent with the behavior of IOs around an elliptical island presented in LH70 and in contrast to coastal inhibition.
The work of Longuet-Higgins provides an initial, though limited, framework to understand the behavior of wind-generated NIOs around island topography similar to Palau. LH69 found the theoretical island-trapped modes that could be excited by a wind forcing around an island. This solution is valid for frequencies equal to f if the island radius is precisely equal to the critical radius, as defined by LH69. LH70 describes linear solutions for IOs near small islands, enabling a simplification of the governing equations by eliminating free waves from the dynamics and only considering motions precisely at f. As such, the solutions are limited in their applicability to the ocean where forcings are broadband, and stratification results in a smaller internal Rossby radius. Furthermore, neither paper considers forcing mechanisms, such as the wind, that excite the trapped or near-inertial motions.
Here, we revisit the work of LH70 and explore the assumptions, dynamics, and energetics of the analytical solution to gain theoretical understanding of IOs around circular and elliptical islands in an infinite domain. Then, to relax some of the constraints of the analytic solutions, we use a wind-forced 1.5-layer reduced-gravity numerical model on an f plane. For simplicity the model dynamics are linear and inviscid. A spatially uniform wind forcing is used to excite NIOs; the periodic boundary conditions we use would admit spatially uniform solutions in the absence of an island. Last, we use a wind patch east of the island to excite near-inertial Poincaré waves (i.e., NIWs) that are partly reflected and partly diffracted as they encounter the island.
This paper is organized as follows: Section 2 explores the analytical solution of rotary, inertial currents around a circular island and an elliptical island, section 3 presents results from a 1.5-layer reduced-gravity model, section 4 summarizes the results and provides concluding remarks.
2. Analytical solution
a. Rotary flow around a circular island forced by boundary oscillations
The pressure pattern in Fig. 1 superficially resembles an azimuthal mode-1 island-trapped wave, but the approximation (6) excludes wave dynamics. The velocity field is horizontally nondivergent, as in the limit of a rigid lid and infinite gravity wave speed.
Despite the similarities of the IO velocity field to that of steady potential flow, the IO and potential flow momentum balances differ, as is evident in the relationship between velocity and the surface elevation (Fig. 1). In steady potential flow, the pressure gradient force balances the advection of momentum. This results in an azimuthal mode-2 structure with pressure maxima at stagnation points around the island, and pressure minima at velocity maxima. For the IO analytical solution, the far-field balance is between the local and Coriolis accelerations, but near the cylinder the balance includes the pressure gradient. At the upstream and downstream stagnation points, local accelerations balance pressure gradients, whereas at the top and bottom edges of the island, the Coriolis acceleration balances the pressure gradient in a Geostrophic balance. These balances also result in an azimuthal mode-2 pattern, but the pressure and velocity extrema now coincide (Fig. 1), instead of being 180° out of phase as in the potential flow example.
The structure of the solution (8), (9), and (10) may also be shown via the normalized polar velocity components and pressure as functions of
b. Rotary flow around an elliptical island
Rotary flow around an elliptical island exhibits many of the same characteristics as flow around a circular island. Regardless of flow direction, there are two stagnation points and two speed maxima on the island boundary. The flow field near the island is largely in geostrophic balance, with high pressure on the left side of the island and low pressure on the right when facing downstream. Once again, the azimuthal wavelength of the surface elevation is equal to the circumference of the island, and the rotation of the far-field velocity forces this pattern to rotate anticyclonically around the ellipse at the inertial frequency.
For an elliptical, rather than circular, island, the rotary flow is blocked by an obstacle of varying shape as the angle of incidence changes. Figure 3 shows the velocity field and surface elevation over a quarter of an inertial period for an ellipse with eccentricity (c = 83) similar to Palau. When the incident flow is exactly normal to the minor axis (Fig. 3a), island blocking and flow acceleration are minimal and the maximum speed is low. A quarter of an inertial period later the incident flow is blocked by the broad side of the island (Fig. 3d), maximizing the flow acceleration around the island tips. Figure 4a shows, for an ellipse of eccentricity c = 83, the azimuthal velocity maximum on the island boundary is approximately 6.4 times greater when the flow is blocked by the major axis than when blocked by the minor axis. The amplitude of the surface elevation on the island boundary is not dependent on the eccentricity (Fig. 4b). The radial pressure gradient force, however, is a function of the eccentricity and increases for a narrower island (larger c), resulting in high flow speeds around the tips of the ellipse (Fig. 4c).
c. Energetics
Kinetic energy is enhanced around the elliptical island tips and reduced along the flanks (Fig. 5b). Additionally, the mean kinetic energy along the major axis of the ellipse is less than in the far field because of the stagnation point that travels along the ellipse boundary. Despite the assumptions of the analytical solution, the distribution of kinetic energy around the ellipse shows many similarities to the distribution of kinetic energy observed around Palau by Siegelman et al. (2019), most notably the observed flow enhancement along the narrow tips of the island chain and weaker inertial currents along the longer island boundaries.
3. Numerical model
In this section, a linear 1.5-layer reduced-gravity model is used to relax the limitations of the analytical solution, providing insight into the more realistic behavior of wind-generated NIOs and NIWs around islands. The model is forced by wind stress, rather than assuming a solution that is periodic at the inertial frequency and that matches an IO in the far field. The idealized wind forcing, with anticyclonic rotation and a Blackman modulation window, is chosen to efficiently excite NIOs. After the forcing has stopped, the flow in the model is free to evolve in time; unlike the analytic solution, it is not constrained to oscillate at a single frequency. In a finite domain and in the presence of a circular island, which introduces a finite length scale, the free solutions to the governing reduced gravity [Eqs. (17)–(19)] consist of Poincaré waves, manifest as NIOs or NIWs for frequencies slightly above f, and island-trapped waves, which resemble Kelvin waves as the island size increases. The model allows us to vary the island size without constraint, to see how the solution evolves as the small-island condition is relaxed. In a final set of numerical experiments, the wind forcing is moved away from the island, showing the response around the island to incident NIWs.
The governing [Eqs. (17)–(19)] are explicitly time integrated using the third-order Adams–Bashforth time-stepping method (Durran 1991). This finite-difference model is solved on a 6000 km × 6000 km, Arakawa C grid with 5-km resolution. Periodic open boundary conditions are used in the far field, with a free-slip condition on the island boundary.
a. Spatially uniform winds
To compare with the LH70 solution, which describes an island in an infinite IO field, a spatially uniform wind forcing is used to excite NIOs. While the theory described in LH70 requires that the island is small relative to the Rossby radius of deformation (ϵL ≪ 1), this condition is removed in numerical simulations where the island radius is varied, but the Rossby radius of deformation is kept constant. In most of the experiments the wind forcing is modulated by an envelope of one inertial cycle. An additional experiment is run for a large-island case with a three-cycle forcing envelope to show the effect of reduced forcing bandwidth.
1) A case study: Dynamics near a small island (ϵL = 0.11)
The motivation for this case study is to assess the behavior of spatially uniform wind-generated NIOs around a dynamically small island after wind forcing has stopped and motion is not constrained to σ = f. This case is compared with LH70’s analytical solution, which requires a small island and imposes spatially uniform IOs. The island radius is 50 km and ϵL = 0.11.
Once the wind forcing stops, NIWs continue to radiate outward, superimposed on the far-field NIOs. Qualitatively, similarities exist between the numerical and analytic solutions of the flow field and surface elevation around a circular island (Fig. 7). As in Fig. 1, the numerical solution has two stagnation points where the flow is normal to the island boundary, and two speed maxima as the flow accelerates around the island after reaching the initial stagnation point. This velocity pattern rotates anticyclonically around the island at approximately the inertial frequency. Furthermore, two surface elevation extrema (one minimum and one maximum) associated with the strongest currents around the island exist.
Unsurprisingly, the analytical and numerical model solutions differ. First, an outward propagating superinertial Poincaré wave is generated at the island boundary even after the wind forcing is shut off. Second, there is a phase lag between the far-field NIO, which is purely zonal at the instant shown in Fig. 7, and the near-field currents and surface elevation. According to the analytical solution (9) and (10), given a far-field, zonal flow, the surface elevation and velocity extrema would be located at the north and south ends of the island (Fig. 1), but here, the extrema have yet to reach these locations. The phase lag is consistent with a slightly subinertial island-trapped wave propagating anticyclonically around the island.
Consistent with the analytical solution, the model is linear and inviscid. Figure 7 shows snapshots of the terms from the numerical model x- and y- momentum balance Eqs. (17) and (18) at one inertial period, just after the wind stops. The far field is primarily a balance between the local acceleration
Quantitative comparisons between the model and analytic solutions of LH70 illustrate some key differences. At one inertial period, as the wind forcing ends, the analytical and modeled solutions are in best agreement (Figs. 8b–d, dashed black line compared to pale red line). Consistent with LH70, the azimuthal velocity at the island boundary is approximately twice that of the far field, and the normalized surface elevation is approximately −2 at the island boundary. Over time, the surface elevation and velocity of the numerical solution amplify, resulting in increasingly larger values compared to the analytical solution (Figs. 8b–d, dashed black line compared to increasingly dark red lines). This enhancement is also seen in the Hovmöller diagrams of these fields along the island boundary (Figs. 9a,c,e). Over time, the amplified region (between the sloping red dashed lines) slowly spreads outward from the island. Beyond this region there is a subtle dip in velocity amplitude (Figs. 9c,e between red and blue dashed lines). After one inertial period there is no wind forcing, so these structures result from the superposition of free waves, including incident and reflected Poincaré waves and the subinertial trapped wave. The energy enhancement will be further discussed in section 3a(2).
The Hovmöller diagram of surface elevation shows slanting phase lines during the entire run corresponding to Poincaré waves radiating away from the island (Fig. 9a). The decreasing slope of the phase lines over time reflects the increase in phase speed of the waves as their frequency approaches, but never reaches, the inertial frequency.
The power spectral density of the surface elevation along the zonal transect emphasizes the island-trapped wave and the energy radiating away from the island (Fig. 9b). Near the island boundary, the spectral peak is wide with a peak frequency at f, indicating energetic sub- and superinertial motions. The subinertial energy is the signature of the island-trapped wave (LH69). Away from the island, consistent with radiating Poincaré waves, the spectral peak is blue shifted (i.e., above f), with little energy near the inertial frequency. The surface elevation expression of NIOs and NIWs vanishes as the frequency approaches f.
Compared to the surface elevation spectrum, the velocity spectrum is more narrow banded and concentrated near f and relatively uniform with distance from the island (Figs. 9d,f). The radial component (u) goes to zero at the boundary over an island radius scale. The variance of the tangential component (υ) increases toward the boundary, consistent with the free slip boundary condition.
2) ϵL dependence
To move away from the small-island case, we increase ϵL by increasing the island radius, a. Immediately after the winds shut off, the phase difference between the far-field NIOs and the currents near the island is greater as ϵL increases (Fig. 10). The angular frequency
The deviation between the numerical results and LH69 is due to the superposition of superinertial, near-inertial, and subinertial motions in the model, while the theory only reflects the subinertial trapped modes (LH69). Evidence of the superposition is seen in the increasing bandwidth of the spectral peak as ϵL increases (Fig. 12). When ϵL = 0.11, the trapped wave frequency is 0.99f, which is consistent with the slight phase lag near the island boundary, while maintaining a single spectral peak (spectral resolution = 0.05 cpd). When ϵL = 1.82, there are two distinct spectral peaks, one at the inertial frequency and the other at the trapped wave frequency, 0.154 cpd. In the far field, there is a slight increase in the variance of η at superinertial frequencies with increasing ϵL, suggesting more energetic Poincaré waves with larger ϵL.
First, consider the small island case, where ϵL = 0.11 (Fig. 13, first row). On the island boundary, energy continues to increase in the absence of wind forcing after one inertial period (also evident in Figs. 9c,d). The negative radial energy flux near the boundary shows that energy from offshore is moving toward the island; energy increases adjacent to the island and decreases farther from the shore. With increasing distance from shore, additional fluctuations in kinetic energy correspond to alternating positive and negative radial energy flux.
The flux of energy within the domain is the result of the changing interference pattern due to the superposition of the incident, radiating and trapped waves. In all simulations, the far field (>1000 km from the island boundary) is dominated by the NIOs and Poincaré waves, which superpose to create regions of increased and decreased kinetic energy that correspond to positive and negative radial energy flux. While the far-field NIOs are identical in all runs, the radial energy flux increases with the island size. A larger island intercepts more near-inertial energy from the large-scale NIO, and therefore reflects more energy away from its boundary in the form of Poincaré waves. A smaller island allows a larger fraction of incident energy to diffract rather than reflect.
To understand the interference pattern near the island (within 1000 km from the island boundary), consider the linear superposition of two waves of differing frequencies (f1 and f2). Due to constructive and destructive interference, the energy fluctuates at the beat frequency, that is, |f1 − f2|. In these simulations, near the island, the two waves are the near-inertial motions and the trapped wave propagating anticyclonically around the island. The near-inertial motions consist of the large-scale NIOs plus the outward-radiating NIWs that are necessary to satisfy the boundary condition at the island. Initially, the energy near the island increases as the trapped wave and the near-inertial motions become more in phase and interfere constructively. When ϵL is small, the trapped wave and the near-inertial motions have nearly the same frequency; thus, the beat period is extremely long and only a small fraction of it occurs during a 20-day run. As a result, for small values of ϵL (0.11, 0.26), less than half the beat period is resolved and only a period of energy increase near the island is seen.
As ϵL increases, the trapped wave frequency decreases and the beat period becomes shorter. The three runs that resolve at least half the beat period (ϵL = 0.46, 1.02, 1.82) show both constructive and destructive interference near the island. As with ϵL = 0.11, energy initially increases on the island boundary. However, after half a beat period, energy decreases on the island boundary and moves offshore, indicated by the positive radial energy flux. Given a long enough run and a large enough island, we see multiple cycles.
3) A large island (ϵL = 4.10): Sensitivity to forcing bandwidth
The previous section demonstrates that the superposition of the large-scale NIOs, radiating Poincaré waves, and subinertial island-trapped waves accounts for the net ocean response near an island to an episode of spatially uniform near-inertial wind forcing. For small islands the trapped wave frequency is too close to f for simple data analysis techniques to isolate that component of the motion in a numerical run lasting a few inertial cycles. The mode-1 trapped wave frequency decreases with increasing island size, though, so in this section we use a large enough island (ϵL = 4.10) to easily distinguish the trapped wave from the remainder of the motion. Additionally, we use a second run with a longer forcing interval, hence a narrower forcing bandwidth (Fig. 14), to show the sensitivity of the response to this parameter. In this run the amplitude is reduced by a third and the modulation envelope spans 3 inertial periods, so the work done by the wind in forcing NIOs is unchanged. For both runs the island radius is 300 km, the domain is expanded to 25 000 km × 25 000 km, and the resolution is reduced to 10 km.
To isolate the island-trapped wave, the time series at each point is bandpass filtered via complex demodulation at the theoretical island-trapped wave frequency, 0.11 cpd, from LH69. The first and last trapped wave periods are subject to filtering end effects and are therefore omitted from the analysis. The residual elevation shows the expected radiating Poincaré wave signal (Fig. 15a), while the residual velocity is dominated by NIO energy that is suppressed near the island (Figs. 15c,e). The trapped wave (Fig. 15, right column) has a Kelvin wave character with energy in the surface elevation and the alongshore velocity component (υ in this zonal transect). The modeled surface elevation as a function of distance from the island agrees well with the LH69 theory, and both are close to the exponential behavior of a Kelvin wave (Fig. S1 in the online supplemental material).
The sensitivity of the NIOs and Poincaré waves to the bandwidth of the wind stress is reflected in comparisons between the 1 and 3 inertial period forcing (1IPF and 3IPF, respectively) runs 3000 km east of the island boundary, beyond the radial extent of the trapped wave. The rotary spectra show that NIOs dominate the far field with anticyclonic, inertial peaks of identical magnitude in each of the numerical runs (Fig. 16a). In the time domain, the inertial peaks manifest as the plateau of kinetic energy at 0.002 m2 s−2 for both runs (Fig. 17a). The kinetic energy plateau is disrupted by the arrival of dispersive Poincaré waves generated at the island boundary. The superposition of these waves on the background NIO field causes fluctuations of the total kinetic energy that correspond to the positive and negative radial energy flux patterns observed in the far field of Fig. 13. Compared to the 3IPF run, the broader forcing bandwidth in the 1IPF run generates higher-frequency waves that travel faster and arrive sooner (Fig. 17a, magenta arrow). The power spectral density of η in the far field reflects only the Poincaré waves because the background NIOs have negligible surface expression. Increasing the forcing period from 1 to 3 inertial periods slightly reduces the near-inertial spectral peak height and decreases its bandwidth. (Fig. 16c).
At the eastern most point on the island boundary, the currents are rectilinear (alongshore), and the clockwise and counterclockwise rotary components are identical (Fig. 16b, dashed line overlies the solid line). The two distinct peaks are at the trapped wave frequency ftw (0.11 cpd, consistent with LH69) and the inertial frequency. Once again, the power spectral density (PSD) at the inertial frequency is identical in the two runs (seen also in the time domain, Fig. 17b). The near-inertial peak is much lower than it is in the far field, consistent with the reduction in residual current near the island seen in Figs. 15c and 15e.
The dynamics driving the attenuation of near-inertial currents toward the island are similar to those along an infinite, straight coastal boundary, known as coastal inhibition (Kundu et al. 1983). After high-frequency Poincaré waves quickly radiate away, the more slowly propagating NIWs remain. These NIWs superpose with the NIOs reducing the amplitude of the near-inertial motions near the island. The residual of the tangential velocity reflects the superposition of the NIWs and NIOs (Fig. 17b). One important difference in the work of Kundu et al. (1983), is that their continuously stratified model allows for downward propagation of near-inertial energy as opposed to our model that only allows for horizontal energy propagation. Despite this difference, both models admit near-inertial waves, which reduce energy in the NIOs near the coastal boundary.
As expected, the PSD in the trapped wave band is lower in the 3IPF run (Figs. 16b,d) because the forcing contains less energy at subinertial frequencies (Fig. 14). The amplitude of the tangential velocity does not decay over time (Fig. 17c), confirming that the energy at this frequency is in fact trapped. (For a view of the spectra as functions of distance from the coast see Fig. S2.)
b. Nonuniform winds
Although comparable to the LH70 analytical solution, the applicability of the results from the previous section is limited by the assumption of uniform wind stress across the entire domain. The generation of subinertial island-trapped waves by local winds has been well documented (Pizarro and Shaffer 1998; Brink 1999; Merrifield et al. 2002), but the excitation of slightly superinertial oscillations by NIWs impinging on an island has not been previously considered. A related problem is the model of Dale et al. (2001) for a straight coast with continuous stratification, a continental shelf, and a spur. They found that a superinertial current along the coast caused radiation from the spur of leaky coastally trapped waves at the same frequency. Similarly, we might expect NIWs impinging on an island to excite leaky island-trapped waves—that is, motions with structure very similar to pure island-trapped modes, but at slightly higher frequency that gradually lose energy by radiating NIWs. Therefore, we conduct simulations in which there is no local wind forcing, but a remote wind patch forces a spectrum of radiating Poincaré waves that impinge on the island. In these simulations, subinertial island-trapped waves cannot be excited because only energy from superinertial Poincaré waves reaches the island.
To examine this case, the original wind forcing, still varying in time as in Fig. 6a, is multiplied by a spatial Blackman window (21) that is nonzero for 1000 < x < 2000 km (blue rectangle in Fig. 18c). The domain extends farther in the east–west direction than the north–south direction (8000 km × 6000 km, 5-km resolution) to increase the distance between the island and the waves that propagate eastward from the source region and reenter in the west through the periodic boundary. Time periods contaminated by this energy reentry are excluded from the analysis.
The zonal structure in the wind excites Poincaré waves that propagate eastward and westward, with the higher wavenumber (higher frequency) components spreading the fastest (Figs. 18a,b). A slowly decaying NIO remains at the axis of symmetry, x = 1500 km. The sloping phase lines in the Hovmöller diagram of surface elevation also show the radiation of Poincaré waves from the axis of wind forcing, with frequency declining toward f over time (Fig. 18c).
The first waves to reach the island are high frequency and short wavelength (Fig. 18). Generally, these waves pass by the island unmodified. As the frequency of the waves decreases toward f and the wavelength increases, a leaky trapped wave is excited that propagates anticyclonically around the island (Figs. 18a,b). The wave has mode-1 structure and a near-inertial peak frequency.
Snapshots for these simulations show NIWs reflected and diffracted by the island superimposed on low wavenumber energy incident from the east and higher-wavenumber energy that has passed the island (Figs. 19a–c). In addition, leaky trapped waves are evident. Their azimuthal modal structure is dependent on ϵL. When the island size is dynamically small (ϵL = 0.11, r = 50 km), the NIWs excite a mode-1 leaky trapped wave. When ϵL = 1.02, a mode-2 wave almost develops, but does not maintain its structure. Finally, when ϵL = 1.82, a mode-2 leaky trapped wave is clearly excited. When the wind is spatially uniform (section 3a), the surface elevation associated with the adjustment around the island projects onto azimuthal mode 1. Here, however, the spatial structure of the incident NIWs can project onto higher modes.
The rotary spectra of the currents at the easternmost point on the island boundary have a broad, slightly blue shifted near-inertial peak (Figs. 19d–f). The wider near-inertial peak, which is higher when ϵL is small, reflects the bandwidth of the NIWs impinging on the island. The total variance of the currents is 3 times larger when ϵL = 0.11 than ϵL = 1.82, indicating stronger near-inertial currents occur near a dynamically small island, which is consistent with the uniform wind forcing runs.
Similar to the velocity spectrum, the spectrum of η on the eastern boundary of the island also shows one slightly blue shifted, near-inertial peak regardless of ϵL (Figs. 19d–f). When ϵL = 1.82, this result may seem surprising because of the distinct subinertial trapped wave peak that arises in response to a spatially uniform wind forcing (Fig. 12c); however, this result is consistent with the theory of LH69. As noted in the introduction, LH69 found that a trapped wave could exist at the inertial frequency under the condition that the island radius is precisely equal to the critical radius [n(n − 1)g′H]1/2/f. Given a Rossby radius of deformation of 148.2 km, the critical radius is 209 km when n = 2. The island radius is 200 km when ϵL = 1.82, so the incident NIWs excite a mode-2 leaky trapped wave. Note the slight blue shift of the near-inertial peak (Fig. 19f), consistent with the fact that only slightly superinertial energy can reach the island in this model. When the island radius is too far from the critical radius for a mode, for example, the radius is 150 km (ϵL = 1.02), a trapped wave is not excited.
4. Summary and conclusions
Results from the 1.5-layer reduced-gravity model show how an episode of spatially uniform wind forcing can generate far-field NIOs, island-trapped waves, and NIWs radiating outward from the island. Similarly, when Poincaré waves generated by a remote wind forcing encounter an island, the dominant response is reflected NIWs, diffraction of incident wave fronts, and the generation of leaky island-trapped waves. In the small-island case (ϵL ≪ 1) with very large-scale forcing, the superposition of the incident, reflected, and trapped waves results in locally enhanced near-inertial motions similar to the analytical solution (section 2). Despite this similarity, there are fundamental differences between the analytic solution and the numerical experiments. The analytic solution, by design, is in a straitjacket: it requires that all motion in the infinite domain is harmonic at a single frequency. That frequency can only be f, since the far-field solution must be an inertial oscillation—the limit of a Poincaré wave as its horizontal scale goes to infinity. In contrast, after the forcing has been turned off, the numerical experiments can be viewed as initial value problems with the initial state being whatever fields of velocity and elevation have been generated by the forcing. Thereafter, the solution is simply the superposition of free waves, which can be island trapped (at discrete frequencies) or propagating to infinity (with a spectrum of superinertial frequencies).
As the island size increases from zero, the mode-1 trapped wave frequency decreases from f (LH69). For ϵL < ∼1 the frequency is close to f, so the relative phases of the trapped wave and the NIWs (i.e., Poincaré waves) change only slowly. With a much larger island, for example, ϵL = 4.1, the frequency difference is obvious, and we can easily separate the trapped wave, with its discrete subinertial frequency, from all the NIWs.
The large-island case differs from the small-island in another important way: the net effect of the large island is to reduce the near-inertial energy near the island, but the inertial frequency band, which includes the slightly subinertial trapped wave, is enhanced near the small island. As ϵL increases, the island reflects more near-inertial energy. As a result, the dominant near-inertial response to the island is not diffraction, but rather reflection. Large amplitude NIWs radiate energy away from the island. What remains locally is a subinertial island-trapped wave and weak near-inertial motions. In the limit of an infinite coastal wall, the island-trapped wave would be replaced by a Kelvin wave and the phenomenon of weak near-inertial motions could be described as coastal inhibition as in the stratified model of Kundu et al. (1983).
The present study has moved beyond the LH70 analytical solution for IOs around idealized island topography by relaxing the small-island limit and by looking at the response to an initial interval of forcing. Additional steps toward understanding near-inertial motion around islands such as Palau include considering more realistic topography and continuous stratification. Brink (2021) found nearly trapped superinertial waves at sloping island topography. While the topography around Palau is steep, and nearly vertical in some locations, it would be interesting to consider how the solution changes once a slope is introduced. Additionally, the influence of nonlinearity and viscosity should be explored to examine generation of vertical vorticity by NIOs shearing against topography. Siegelman et al. (2019) observed a period of enhanced near-inertial currents that corresponded to a large scale wind event near the narrow, northern tip of Velasco Reef, Palau. During this period, the vorticity estimated from a mooring array deployed near the reef boundary had energy in the near-inertial band. While Siegelman et al. (2019) hypothesized that near-inertial vorticity was generated due to enhanced NIOs shearing against the reef wall, this has not yet been supported by a numerical or theoretical study.
Acknowledgments.
We thank the two anonymous peer reviewers for their thoughtful comments. We acknowledge Brian Powell for helping to lay the foundation for the numerical modelling used in this paper. Pat and Lori Colin at the Coral Reef Research Foundation were essential to the fieldwork campaign at Palau, which motivated our efforts. This work was funded by the Office of Naval Research through Grant N00014-16-1-2671. Mika Siegelman was partially funded by the Denise B. Evans Fellowship, and Ruth C. Musgrave was supported, in part, thanks to funding from the Canada Research Chairs Program.
Data availability statement.
The model output analyzed in this paper is available from https://doi.org/10.6075/j0j67h3p.
APPENDIX A
Nondimensionalization of the Long-Wave Equation
APPENDIX B
Analytical Solution
Expanding on sections 2a and 2b, we present the detailed solution to the analytical work of LH70. We revisit this work to correct errors and clarify steps in his original publication.
a. Circular island
b. Elliptical island
REFERENCES
Brink, K. H., 1999: Island-trapped waves, with application to observations of Bermuda. Dyn. Atmos. Oceans, 29, 93–118, https://doi.org/10.1016/S0377-0265(99)00003-2.
Brink, K. H., 2021: Near-resonances of superinertial and tidal fluctuations at islands. J. Phys. Oceanogr., 51, 2721–2733, https://doi.org/10.1175/JPO-D-20-0298.1.
Dale, A. C., J. M. Huthnance, and T. J. Sherwin, 2001: Coastal-trapped waves and tides at near-inertial frequencies. J. Phys. Oceanogr., 31, 2958–2970, https://doi.org/10.1175/1520-0485(2001)031<2958:CTWATA>2.0.CO;2.
D’Asaro, E. A., 1989: The decay of wind-forced mixed layer inertial oscillations due to the β effect. J. Geophys. Res., 94, 2045–2056, https://doi.org/10.1029/JC094iC02p02045.
Durran, D. R., 1991: The third-order Adams-Bashforth method: An attractive alternative to leapfrog time differencing. Mon. Wea. Rev., 119, 702–720, https://doi.org/10.1175/1520-0493(1991)119<0702:TTOABM>2.0.CO;2.
Fontán, A., and B. Cornuelle, 2015: Anisotropic response of surface circulation to wind forcing, as inferred from high-frequency radar currents in the southeastern Bay of Biscay. J. Geophys. Res. Oceans, 120, 2945–2957, https://doi.org/10.1002/2014JC010671.
Kundu, P. K., S.-Y. Chao, and J. P. McCreary, 1983: Transient coastal currents and inertio-gravity waves. Deep-Sea Res., 30A, 1059–1082, https://doi.org/10.1016/0198-0149(83)90061-4.
Longuet-Higgins, M. S., 1969: On the trapping of long-period waves round islands. J. Fluid Mech., 37, 773–784, https://doi.org/10.1017/S0022112069000875.
Longuet-Higgins, M. S., 1970: Steady currents induced by oscillations round islands. J. Fluid Mech., 42, 701–720, https://doi.org/10.1017/S0022112070001568.
Merrifield, M. A., L. Yang, and D. S. Luther, 2002: Numerical simulations of a storm-generated island-trapped wave event at the Hawaiian Islands. J. Geophys. Res., 107, 3169, https://doi.org/10.1029/2001JC001134.
Pizarro, O., and G. Shaffer, 1998: Wind-driven, coastal-trapped waves off the island of Gotland, Baltic Sea. J. Phys. Oceanogr., 28, 2117–2129, https://doi.org/10.1175/1520-0485(1998)028<2117:WDCTWO>2.0.CO;2.
Pollard, R. T., 1980: Properties of near-surface inertial oscillations. J. Phys. Oceanogr., 10, 385–398, https://doi.org/10.1175/1520-0485(1980)010<0385:PONSIO>2.0.CO;2.
Pollard, R. T., and R. C. Millard, 1970: Comparison between observed and simulated wind-generated inertial oscillations. Deep-Sea Res. Oceanogr. Abstr., 17, 813–821, https://doi.org/10.1016/0011-7471(70)90043-4.
Schahinger, R. B., 1988: Near-inertial motion on the South Australian shelf. J. Phys. Oceanogr., 18, 492–504, https://doi.org/10.1175/1520-0485(1988)018<0492:NIMOTS>2.0.CO;2.
Shearman, R. K., 2005: Observations of near-inertial current variability on the New England shelf. J. Geophys. Res., 110, C02012, https://doi.org/10.1029/2004JC002341.
Siegelman, M., and Coauthors, 2019: Observations of near-inertial surface currents at Palau. Oceanography, 32, 74–83, https://doi.org/10.5670/oceanog.2019.413.
Tintoré, J., D.-P. Wang, E. Garćia, and A. Viúdez, 1995: Near-inertial motions in the coastal ocean. J. Mar. Syst., 6, 301–312, https://doi.org/10.1016/0924-7963(94)00030-F.