1. Introduction
There is evidence that planetary-scale Rossby waves have been generated off the western coast of the United States, either by unstable coastal boundary currents or by coastal waves matching in frequency and scale (cf. Kelly et al. 1998). Bokhove and Johnson (1999), therefore, investigated the matching of planetary Rossby modes with coastal shelf modes in a cylindrical basin. Otherwise said, linear free modes were calculated with so-called semianalytical “mode matching” techniques, as well as linear forced–dissipative finite-element methods, to find resonances. Two parameter regimes were considered: an ocean one and a laboratory analog. These laboratory-scale hybrid Rossby-shelf modes had been considered with validating laboratory rotating tank experiments in mind.
Such an experimental validation is the topic of the present paper. Planetary barotropic Rossby modes have been shown before in the laboratory using the analogy between planetary β–plane Rossby modes and topographic shelf modes for a uniform basin-scale north–south background topography—for example, in the classic book of Greenspan (1968). Rotating tank experiments geared toward enforcing resonant hybrid coastal and planetary modes appear to be (relatively) new. In preparation of the rotating tank experiments, linear forced–dissipative finite-element calculations of barotropic potential vorticity dynamics have revealed the resonant frequencies of two primary hybrid Rossby-shelf modes. These primary forcing frequencies were then used to drive the harmonic “wind” forcing provided via the Ekman pumping and suction due to an oscillating rigid lid. Various forcing strengths have been imposed in which a match with the linear calculations is best suited by weak forcing, whereas better visualization requires a larger signal-to-noise ratio and, consequently, stronger forcing. The latter promotes, however, the emergence of nonlinear effects. We, therefore, also compare the experimental streamfunction fields with some nonlinear simulations of barotropic potential vorticity dynamics in an attempt to explain the differences between linear theory and the experimental results under stronger forcing.
In ocean general circulation models (OGCMs), the coastal regions are often underresolved; furthermore, it is hypothesized that significant energy exchange takes place between the deeper ocean and the shallower coastal zones, for example, Wunsch (2004). Vertical walls placed in the shallow seas are generally used in coastal zones as lateral boundaries in OGCMs, whereas in reality, mass, momentum, and energy are transferred across these virtual vertical walls. As a consequence, such an exchange would not be resolved or modeled properly in the OGCMs. Suitable parameterizations of these underresolved processes would therefore be required in the coastal zones of OGCMs. The laboratory experiments and finite element calculations we present aim to serve as an idealized barotropic system to investigate this modal coupling between basin- and coastal-scale dynamics.
The outline of this paper is as follows: barotropic potential vorticity dynamics is introduced in section 2, and linear finite-element calculations are presented to find the relevant forcing frequencies. These forcing frequencies are a building block in section 3, where the experimental setup and results are presented. Preliminary nonlinear simulations in section 4 indicate the effects of strong forcing on the dynamics observed. A short conclusion is found in section 5.
2. Rigid-lid potential vorticity model
a. Nonlinear model






















The numerical (dis)continuous Galerkin finite-element discretization used is based on formulation (6), by extending the inviscid formulation in Bernsen et al. (2006); it couples the hyperbolic potential vorticity Eq. (6a) to the elliptic Eq. (6c) for the transport streamfunction and is advantageous for complex-shaped domains. Instead of a classical continuous finite-element method as used for the streamfunction, potential vorticity is descretized discontinuously. For smooth profiles of potential vorticity, the numerical discontinuities between elements are negligible and scale with the mesh size and order of accuracy. The numerical method conserves vorticity and energy for infinitesimally small time steps in the inviscid and unforced case, whereas enstrophy is slightly decaying for the upwind flux used. The weak formulation of this finite-element method is given in the appendix. The method provides an alternative to classical numerical methods and is well suited for complex-shaped domains, mesh, and order (h and p) refinement.
b. Linear model and hybrid Rossby-shelf modes
We have performed laboratory experiments to assess whether hybrid Rossby-shelf modes exist that couple planetary-scale Rossby modes with coastal-scale shelf modes. In the laboratory experiments, a uniformly rotating tank is used with rotation frequency Ωr. Consequently, there is no planetary variation of the background rotation in the north–south direction. We consider the Northern Hemisphere and define a north–south direction indirectly by introducing a background slope s = s(y) in the y direction, which is related mathematically to the β effect (e.g., Greenspan 1968). Hence, the nondimensional β parameter β = s/


The forced–dissipative response of the laboratory ocean, described above and sketched in Fig. 1, is displayed in Fig. 2 for κ =
In the nonlinear numerical simulations described in the next section, instead of the abrupt step shelf break, a smoothed shelf break is introduced between Rs − ϵ < r < Rs + ϵ. The finite element mesh contains regular nodes placed at the circles with radii Rs ± ϵ and R; then random nodes are added, subject to a minimum distance criterion outside the shelf break; subsequent triangulation yields a triangular mesh; and additional nodes are placed at the centroid and midpoints of the edges of each triangle to further divide the triangular mesh into a quadrilateral one. The shelf break contains two elements across and upon mesh refinement, the value of ϵ decreases and hence the shelf break also becomes narrower. For such smoothed topography, the linear forced–dissipative response and the streamfunction field at the maximum resonance are given in Figs. 6 and 7 for ϵ = 0.0314. Relative to the abrupt shelf topography with resonant forcing frequency σ = 0.0613, the resonant frequency in the new calculation of the linearized system has lowered by 6% to σ = 0.0577, whereas the actual fields at resonance remain highly similar.
3. Laboratory experiments
a. Experimental setup


b. Coupled modes
Numerous experiments were carried out for a few forcing frequencies. For each experiment, streak photography was obtained with 2–4-s exposure time. The main forcing frequencies used in the laboratory experiments were calculated with a finite-element model of the linearized equations, as explained in section 2b. We report here solely four sets of experiments, deemed best for their visual resolution and the hybrid character of the Rossby-shelf mode.
The two sets of eight images in Figs. 9 and 10 give an impression of the flow during one forcing period of 51.3 s—that is, with dimensional frequency σ* = 0.1226 s−1 (σ = 0.0612). It nearly corresponds to a numerically calculated forced–dissipative resonance for a hybrid Rossby-shelf mode of the rigid-lid model (6), linearized around a state of rest (see Figs. 2 and 3). The underlying Rossby mode has azimuthal mode number zero, whereas the underlying shelf mode has azimuthal number m = 1, 2. The shelf break is visible as a thin whitish line at r = 0.8R. In the time sequence from top left to bottom right, a Rossby mode circulation cell in the deep interior “ocean” travels westward, where it absorbs onto the shelf and propagates counterclockwise (in the Northern Hemisphere) as a trapped shelf mode circulation cell. On the eastern boundary, this shelf mode radiates into a planetary Rossby mode. Apart from the striking qualitative resemblance with linear forced–dissipative calculations in Fig. 3, discrepancies occur in the northwest, presumably as a result of nonlinear effects, and in the east where the shelf mode disappears, presumably as a result of strong damping. The rigid lid or glass plate has rotated from zero to a relatively large angle, 2π and π (Figs. 10 and 9, respectively), and back during a period. The nonlinear oscillations in the northwest corner at t = 14, 42 s in Fig. 10 are larger under the greater forcing amplitude than at t = 0, 28 s in Fig. 9. These oscillations diminish even more when the forcing amplitude is reduced to π/2, which is not shown here. The experimental dilemma is that a comparision with linearized modal solutions requires a weak forcing, whereas good visualization requires strong forcing for the streak photography used. Particle image velocimetry techniques could have been used for weak forcing as well, but they were not available in 1997 at the rotating tank facilities in Woods Hole. Although the tank dimensions are not shallow, the simplifying assumption has been that the rotation is sufficiently strong to render the flow to be nearly two-dimensional outside the thin Ekman top and bottom boundary layers, the sidewall boundary layers, as well as the internal and boundary layers at the shelf break.
To compare the amplitudes observed and calculated, the streaks under 4-s exposure are compared with streak lengths in the calculation for σ = 0.006 13 and Δθ = π in Fig. 11. Note that the calculated streaks are weaker, about 40%, than the observed streaks. Such a difference also occurs for forcing with Δθ = 2π in Fig. 12. The precise location of the mode around resonance, and hence its amplitude, as well as nonlinear shifts, might cause this discrepancy. The dimensionless speed at (x, y) = (−0.11, 0.11) is about 0.0276; at the southern shelf, the maximum speed is about 0.0773 in Fig. 11a. The speed at (x, y) = (0, 0) is about 0.0543; at the southern shelf, the maximum speed is about 0.1087 in Fig. 12a.
Similarly, the linear forced–dissipative mode calculated at σ = 0.0878 corresponds reasonably well with the observed mode for σ = 0.0871 (based on the FEM calculation in 1997) in Fig. 13 concerning observations for stronger forcing with Δθ = 2π.
Finally, we conclude that hybrid Rossby-shelf modes exist and can be successfully visualized and measured in the laboratory; they correspond well with linear forced–dissipative calculations at a similar frequency. Discrepancies between the experimental and numerical flow patterns are observed especially at the northwestern boundary. Additional nonlinear simulations aim to explain this discrepancy.
4. Laboratory results versus numerical simulations
Nonlinear simulations at resonance frequency σ = 0.0613 have been performed, starting from a state of rest and with sinusoidal forcing. The forcing period is thus 102.5 time units; in addition, κ = 0.0042, with a smoothed shelf break of width 2ϵ = 0.0628, and Δθ = 2π. The value of κ = 0.0042 implies that start-up transients disappear below 1% of their initial value within about 11 periods (cf. the energy and enstrophy graphs versus time in Fig. 14). We note that the solution appears quasi periodic for t > 800. Shown is the solution over period 20 in Fig. 15 (i.e., from t = 1947.5 to 2050.0). A second-order spatial and third-order temporal discretization has been used; single- and double-resolution runs have been performed with 4671 and 10 363 nodes and 4590 and 10 242 elements; the former are shown but agree well with the latter. These nonlinear simulations reveal the cause of the disturbances in the northwest corner of the domain at t = 14, 42 s in Fig. 10: a vortex starts to roll up on the northern shelf once the cell of the southern shelf mode starts to radiate into a basin Rossby mode; subsequently, the vortex gets advected counterclockwise around the domain by the basin Rossby mode and is dissipated once the new forcing cycle starts (see Figs. 15 and 16 in tandem). The simulated potential vorticity field displayed is less smooth than the streamfunction fields and displays more structure, including a vortex shedding.
5. Conclusions
Hybrid Rossby–shelf modes were shown to exist analytically and numerically in Bokhove and Johnson (1999). Based on depth-averaged potential vorticity dynamics, we showed numerically that these hybrid modes also emerged as linear forced–dissipative solutions on a laboratory scale. These hybrid modes matched the largest planetary-scale Rossby mode to a trapped shelf mode—the latter propagating around the southern shelf. The calculated frequencies of the two dominant hybrid modes were used as driving frequency in laboratory tank experiments, based on the linearized calculations. Therein a driven lid provided the Ekman forcing. The spatial structure of the streamfunction fields in the linear calculations and the observed streamfunction fields in streak photography agreed well or reasonably well in the weaker and stronger forcing cases. Discrepancies in amplitude and structure were attributed to nonlinear effects, and nonlinear simulations of the depth-averaged flow suggested observed differences to be due to a vortex generated and shed off the northern shelf by the large westward-propagating planetary Rossby mode in the deep basin. The minimum and maximum amplitudes in the calculations of the linear and nonlinear models differed: we observed values of −0.005 and 0.009 in the former and values of −0.014 and 0.005 in the latter. The simulations provided extra information on potential vorticity dynamics, which was unavailable from the laboratory observations.
Even though the topography used is still simple, the exhibited mode merging shows that the linear normal modes rapidly obtain a complicated structure. The distinction between trapped shelf modes and planetary modes becomes less clear in complex domains and is a bit artificial as both modes emerge from the spatial structure in the background potential vorticity. Vortical normal modes have recently been used to explain temporal variability in the Mascarene Basin (Weijer 2008) and in the Norwegian and Greenland gyres (LaCasce et al. 2008). Unexplored yet interesting aspects in the idealized ocean basin used by us concern the effects of a midocean ridge (cf. Pedlosky 1996) on the communication between two separate deep-ocean half basins connected only via coastal shelves and the numerical parameterization of underresolved shelf mode dynamics on the deep-ocean dynamics as a way to explore energy exchange through an effective, permeable boundary.
Acknowledgments
It is a great pleasure to acknowledge the assistance of John Salzig in the laboratory. Without his help, the experiments would have failed. The laboratory experiments were performed while O. B. was a postdoctoral scholar at the Woods Hole Oceanographic Institution (1996–97); preliminary results were posted in Bokhove (1999). The encouragements and advice of Karl Helfrich, Joe Pedlosky, and Jack Whitehead have, as always, been of great value. Jack Whitehead also kindly allowed O. B. to use the rotating table in his laboratory, and additional support came via Joe Pedlosky’s NSF Grant OCE–9901654 in 1997.
REFERENCES
Bernsen, E., O. Bokhove, and J. van der Vegt, 2006: A (dis)continuous finite element model for generalized 2D vorticity dynamics. J. Comput. Phys., 211 , 719–747.
Bokhove, O., 1999: Forced-dissipative response for coupled planetary Rossby and topographic shelf modes in homogeneous, cylindrical oceans. Preprints, 12th Conf. on Atmospheric and Oceanic Fluid Dynamics, New York, NY, Amer. Meteor. Soc., 104–107. [Available online at http://ams.confex.com/ams/older/aofd12/abstracts/245.htm].
Bokhove, O., and E. R. Johnson, 1999: Hybrid coastal and interior modes for two-dimensional flow in a cylindrical ocean. J. Phys. Oceanogr., 29 , 93–118.
Cenedese, C., J. Whitehead, J. Pedlosky, and S. Lentz, 2007: 2007 Program of studies: Boundary layers. Tech. Rep. WHOI-2008-05, WHOI, 319 pp. [Available online at https://darchive.mblwhoilibrary.org/handle/1912/2503].
Greenspan, H., 1968: The Theory of Rotating Fluids. Cambridge University Press, 354 pp.
Kelly, K., R. Beardsley, R. Limeburner, K. Brink, J. Paduan, and T. Chereskin, 1998: Variability of the near-surface eddy kinetic energy in the California Current based on altimetric, drifter, and moored current data. J. Geophys. Res., 103 , 13067–13083.
LaCasce, J., O. Nost, and P. Isachsen, 2008: Asymmetry of free circulations in closed ocean gyres. J. Phys. Oceanogr., 38 , 517–526.
Pedlosky, J., 1987: Geophysical Fluid Dynamics. 2nd ed. Springer-Verlag, 710 pp.
Pedlosky, J., 1996: Ocean Circulation Theory. Springer, 453 pp.
Weijer, W., 2008: Normal modes of the Mascarene Basin. Deep-Sea Res. I, 55 , 128–136.
Wunsch, C., 2004: Vertical mixing, energy, and the general circulation. Annu. Rev. Fluid. Mech., 36 , 281–314.
APPENDIX
(Dis)continuous Galerkin Finite Element Discretization
Weak formulation








In Bernsen et al. (2006), a finite-element discretization is given and verified for the inviscid, unforced version of (A1) for complex-shaped, multiconnected domains. The generalized streamfunction and vorticity formulation (A1) is advantageous because it unifies several systems into one, such as the barotropic quasigeostrophic, and rigid-lid equations. A third-order Runge–Kutta discretization in time and second-, third-, or fourth-order discretizations in space are implemented and available for use. Without forcing and dissipation, discrete energy conservation is guaranteed in space, whereas the discrete enstrophy decays for the upwind numerical flux and is conserved for the central flux for infinitesimal time steps. The latter central flux is stable but yields small oscillations in combination with the third-order time integrator. When necessary, the circulation along the boundary can also be treated properly on the discrete level.






Normal mode numerical tests
Linear free and forced–dissipative planetary Rossby modes






Linear free shelf mode




























Forced–dissipative hybrid-shelf modes



Sketch of laboratory domain with abrupt shelf topography, and deep interior ocean and shallow-shelf slopes mimicking β (Bokhove and Johnson 1999).
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Sketch of laboratory domain with abrupt shelf topography, and deep interior ocean and shallow-shelf slopes mimicking β (Bokhove and Johnson 1999).
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1
Sketch of laboratory domain with abrupt shelf topography, and deep interior ocean and shallow-shelf slopes mimicking β (Bokhove and Johnson 1999).
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Linear forced–dissipative response for topographic Rossby-shelf modes displayed as the L∞ − norm of |Ψ| against 500 forcing frequencies σ. Parameter values
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Linear forced–dissipative response for topographic Rossby-shelf modes displayed as the L∞ − norm of |Ψ| against 500 forcing frequencies σ. Parameter values
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1
Linear forced–dissipative response for topographic Rossby-shelf modes displayed as the L∞ − norm of |Ψ| against 500 forcing frequencies σ. Parameter values
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Streamfunction field for Rossby-shelf modes over one forcing period T nearby the maximum response σ = 0.0613. Parameter values
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Streamfunction field for Rossby-shelf modes over one forcing period T nearby the maximum response σ = 0.0613. Parameter values
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1
Streamfunction field for Rossby-shelf modes over one forcing period T nearby the maximum response σ = 0.0613. Parameter values
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Streamfunction field for Rossby-shelf modes over one forcing period T at the maximum response σ = 0.0878.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Streamfunction field for Rossby-shelf modes over one forcing period T at the maximum response σ = 0.0878.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1
Streamfunction field for Rossby-shelf modes over one forcing period T at the maximum response σ = 0.0878.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Dispersion relation of the free planetary Rossby mode of zeroth order in the radial direction; the coastal shelf mode for β = 0.3125; H1 = 0.6R, H2 = 0.8R, and Rs = 0.8R.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Dispersion relation of the free planetary Rossby mode of zeroth order in the radial direction; the coastal shelf mode for β = 0.3125; H1 = 0.6R, H2 = 0.8R, and Rs = 0.8R.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1
Dispersion relation of the free planetary Rossby mode of zeroth order in the radial direction; the coastal shelf mode for β = 0.3125; H1 = 0.6R, H2 = 0.8R, and Rs = 0.8R.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Linear forced–dissipative response for topographic Rossby-shelf modes displayed as the L∞ − norm of |Ψ| against 500 forcing frequencies σ for a smoothed shelf break of width 2ϵ = 0.0628 around r = Rs. Parameter values
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Linear forced–dissipative response for topographic Rossby-shelf modes displayed as the L∞ − norm of |Ψ| against 500 forcing frequencies σ for a smoothed shelf break of width 2ϵ = 0.0628 around r = Rs. Parameter values
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1
Linear forced–dissipative response for topographic Rossby-shelf modes displayed as the L∞ − norm of |Ψ| against 500 forcing frequencies σ for a smoothed shelf break of width 2ϵ = 0.0628 around r = Rs. Parameter values
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Streamfunction field for Rossby–shelf modes over one forcing period T at the maximum response σ = 0.0577 for a smoothed shelf break. Parameter values
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Streamfunction field for Rossby–shelf modes over one forcing period T at the maximum response σ = 0.0577 for a smoothed shelf break. Parameter values
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1
Streamfunction field for Rossby–shelf modes over one forcing period T at the maximum response σ = 0.0577 for a smoothed shelf break. Parameter values
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Sketch of the laboratory setup.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Sketch of the laboratory setup.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1
Sketch of the laboratory setup.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Streak photography of hybrid Rossby-shelf modes at t = 0, 7, 14, 28, 35, 42, and 49 s for a forcing period of 51.3 s (nondimensional σ = 0.0612), and maximum rigid-lid excursion Δθ = π. Exposure time was 4 s.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Streak photography of hybrid Rossby-shelf modes at t = 0, 7, 14, 28, 35, 42, and 49 s for a forcing period of 51.3 s (nondimensional σ = 0.0612), and maximum rigid-lid excursion Δθ = π. Exposure time was 4 s.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1
Streak photography of hybrid Rossby-shelf modes at t = 0, 7, 14, 28, 35, 42, and 49 s for a forcing period of 51.3 s (nondimensional σ = 0.0612), and maximum rigid-lid excursion Δθ = π. Exposure time was 4 s.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Streak photography of hybrid Rossby–shelf modes at t = 0, 7, 14, 28, 35, 42, and 49 s for a forcing period of 51.3 s (nondimensional σ = 0.0612), and maximum rigid-lid excursion Δθ = 2π. Exposure time was 2 s.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Streak photography of hybrid Rossby–shelf modes at t = 0, 7, 14, 28, 35, 42, and 49 s for a forcing period of 51.3 s (nondimensional σ = 0.0612), and maximum rigid-lid excursion Δθ = 2π. Exposure time was 2 s.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1
Streak photography of hybrid Rossby–shelf modes at t = 0, 7, 14, 28, 35, 42, and 49 s for a forcing period of 51.3 s (nondimensional σ = 0.0612), and maximum rigid-lid excursion Δθ = 2π. Exposure time was 2 s.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

(a) Streak photography of hybrid Rossby–shelf mode observed at t = 0 s for a forcing period of 51.3 s (nondimensional σ = 0.0612), and maximum rigid-lid excursion Δθ = π. Exposure time was 4 s. (b) Same as in (a), but for the calculated linear solution; phase shift adjusted semioptimally by eye.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

(a) Streak photography of hybrid Rossby–shelf mode observed at t = 0 s for a forcing period of 51.3 s (nondimensional σ = 0.0612), and maximum rigid-lid excursion Δθ = π. Exposure time was 4 s. (b) Same as in (a), but for the calculated linear solution; phase shift adjusted semioptimally by eye.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1
(a) Streak photography of hybrid Rossby–shelf mode observed at t = 0 s for a forcing period of 51.3 s (nondimensional σ = 0.0612), and maximum rigid-lid excursion Δθ = π. Exposure time was 4 s. (b) Same as in (a), but for the calculated linear solution; phase shift adjusted semioptimally by eye.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Same as in Fig. 11, but with maximum rigid-lid excursion Δθ = 2π, and exposure time was 2 s.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Same as in Fig. 11, but with maximum rigid-lid excursion Δθ = 2π, and exposure time was 2 s.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1
Same as in Fig. 11, but with maximum rigid-lid excursion Δθ = 2π, and exposure time was 2 s.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Streak photography of hybrid Rossby-shelf modes at t = 0, 5, 10, 15, 20, 25, 30, and 35 s for a forcing period of 36.1 s (nondimensional σ = 0.0871), and maximum rigid-lid excursion Δθ = 2π.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Streak photography of hybrid Rossby-shelf modes at t = 0, 5, 10, 15, 20, 25, 30, and 35 s for a forcing period of 36.1 s (nondimensional σ = 0.0871), and maximum rigid-lid excursion Δθ = 2π.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1
Streak photography of hybrid Rossby-shelf modes at t = 0, 5, 10, 15, 20, 25, 30, and 35 s for a forcing period of 36.1 s (nondimensional σ = 0.0871), and maximum rigid-lid excursion Δθ = 2π.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Energy and enstrophy vs time.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Energy and enstrophy vs time.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1
Energy and enstrophy vs time.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Streamfunction over forcing period 20; σ = 0.0613, κ = 0.0042, ϵ = 0.0314, and Δθ = 2π.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Streamfunction over forcing period 20; σ = 0.0613, κ = 0.0042, ϵ = 0.0314, and Δθ = 2π.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1
Streamfunction over forcing period 20; σ = 0.0613, κ = 0.0042, ϵ = 0.0314, and Δθ = 2π.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Same as in Fig. 15, but for potential vorticity.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1

Same as in Fig. 15, but for potential vorticity.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1
Same as in Fig. 15, but for potential vorticity.
Citation: Journal of Physical Oceanography 39, 10; 10.1175/2009JPO4101.1