Tutorial 1.5

Baroclinic wave flows in a two-layer fluid

One of the earliest theoretical models for baroclinic instability, proposed in 1951 by Norman Phillips (Phillips 1951), simplified the vertical density stratification of the basic background flow by assuming two superposed, uniform layers of fluid of densities \(\rho_1\) (upper) and \(\rho_2\) (lower), such that the upper layer density is \(\rho_1 < \rho_2\) in a statically stable configuration – see Fig. 1.5a below. The initial flow consists of uniform zonal motion (in the \(x\)-direction) in each layer, with eastward flow \(u_1=+\frac{U}{2}\) in the upper layer and westward flow \(u_2=-\frac{U}{2}\) in the lower layer, along a straight channel of width \(L\) and total depth \(D\).

Diagram of the two-layer model
Fig. 1.5a. The two-layer model considered here, consisting of two immiscible layers of different densities in relative motion. (Source: Lovegrove et al. 2001)

To maintain this flow in a steady state, in geostrophic balance with the background rotation speed \(\Omega\), the interface between the two layers must therefore slope upwards towards increasing \(y\) (northwards in the Northern Hemisphere). The slope of the interface must scale as \[\frac{\partial h_2}{\partial y} = \frac{2 \Omega}{g'}(u_1-u_2) = \frac{2 \Omega U}{g'}\] where \(g' = g \frac{\rho_2-\rho_1}{\rho_1}\) is the reduced gravity, thereby storing potential energy.

In principle, this potential energy can be released in the same way as in continuously stratified flows (see section 1), by the action of a wave-like baroclinic instability, provided that other dynamical constraints are satisfied. A more detailed model calculation, taking into account Ekman friction and the dependence of the Coriolis parameter \(f\) with latitude \(y\) (so \(f=f_0+\beta y\)), shows that the onset of baroclinic instability (and its nonlinear development) depend on three main dimensionless parameters:

  1. Froude number \[F=\frac{2 f_0^2 L^2}{g' D}\]
  2. Beta parameter \[B=\frac{\beta L^2}{U}\]
  3. Dissipation parameter \[r=\frac{\sqrt{2 v f_0} L}{U D}\]

The overall predicted regime diagram for this model is shown below (from Lovegrove et al. 2002):

Regime diagram for a two-layer model
Fig. 1.5b. Three-dimensional regime diagram for the two-layer model, in the \([r, \beta, F]\) parameter space. The \(\mathrm{O}(2)\) symmetry plane (where \(\beta=0\)) and the inviscid plane (where \(r=0\)) are indicated. (Source: Lovegrove et al. 2002)

This system is not just an abstract mathematical theory, however – it can be realised and explored in laboratory experiments. The figures below show a rotating annulus experiment that uses two immiscible fluids in uniform layers in a cylindrical annular tank. In practice, water was used as the upper layer and a mixture of an organic oil (D-Limonene) with a dense chlorofluorocarbon liquid (CFC-113) used for the lower layer, resulting in two layers with a density difference \(\Delta\rho=6\text{ kg m}^{-3}\), a mean density \(\bar{\rho}=1000\text{ kg m}^{-3}\), and reduced gravity \(g'=5.9\text{ cm s}^{-2}\).

Schematic diagram of the apparatus used in the two-layer annulus experiment
Fig. 1.5c. Schematic cross-section through the apparatus used to investigate the two-layer annulus, showing the principal mechanical components (not to scale). (Source: Williams et al. 2004)
Apparatus for the two-layer annulus in the laboratory
Fig. 1.5d (left) and Fig. 1.5e (right). Photographs of the apparatus for the two-layer annulus in the laboratory, general and close-up.

The differential zonal motion between the upper and lower layers is maintained mechanically by rotating the lid of the tank with respect to the rest of the tank (in either direction). The requirement for geostrophic balance leads to the interface between the layers becoming parabolically sloped in radius, as shown schematically above. When waves develop as a result of instabilities, these appear as wave-like perturbations to the interface in geostrophic balance with horizontal velocity perturbations in each layer.

We visualise these perturbations to the interface using a clever optical method, due originally to Hart & Kittleman 1985. The lower layer includes the material D-Limonene, an organic oil distilled from the peel of citrus fruits such as oranges or lemons. This is strongly optically active, which means it systematically rotates the plane of polarization of by an amount that depends on the optical path length and the wavelength of light. The variation of the angle of rotation at fixed optical path length is shown below, which clearly shows the strong wavelength dependence:

Plot of wavelength against rotation angle, illustrating a strong dependence
Fig. 1.5f. Plot of wavelength against rotation angle, illustrating a strong (negative) dependence. (Source: Lovegrove et al. 2000)

As indicated above, we shine white, plane polarized light upwards through the bottom of the tank through a neutral diffuser, towards a colour video camera located above the tank on the rotation axis. By adding a second polarising filter between the camera and the top of the tank, we allow only a single wavelength of light to come into alignment with the second polarizer for a given optical path length. But if the path length varies across the tank because of dynamical perturbations to the interface, different heights are mapped onto different colours!

Still from a video of a rotating annulus, with different colours indicating different heights
Fig. 1.5g. Still from a video of a rotating two-layer annulus, with different colours indicating different interface heights – a wavenumber \(m=9\) pattern is clearly visible. Blue regions correspond to relatively large values of interface height, and yellow regions to relatively low values. The turntable and lid rotation speeds, \(\Omega\) and \(\Delta\Omega\), respectively, are both positive. Since the video camera that recorded this image is fixed in the turntable frame, the turntable appears to be stationary and the lid (and disturbances) rotate in the anticlockwise sense. (Source: Lovegrove et al. 2000)
Reconstructed interface height
Fig. 1.5h. Reconstructed two-dimensional interface height, as inferred from the image in Fig. 1.5g, by projecting the hue onto the mean of calibration curves. (Source: Lovegrove et al. 2000)

With care, the colour mapping can even be calibrated to recover a quantitative map of the interface, as shown above (from Williams et al. 2004).

This method can then be used to explore what kinds of flow may occur in different regions of parameter space, mainly with respect to the Froude number \(F\) and dissipation parameter \(r\).

Regime diagram with the Froude number against the dissipation parameter
Fig. 1.5i. Regime diagram with the Froude number against the dissipation parameter
Regime diagram showing the location of different observed flow types
Fig. 1.5j. Regime diagram showing the location of the different observed flow types in the experimental \([r,F]\)-plane. Numbers within each region denote the dominant baroclinic wave number, while letters denote the type of flow:
S: steady wave (apart from a slow drift around the annulus)
Av: periodic amplitude vacillation
Sv: structural vacillation
I: intermittent bursting flow
F: anomalous forced-resonant flow
Lists such as 2, 1I, etc. denote multiple equilibria.

Various flow regimes are indicated here, including:

AX: Axisymmetric flow
KH: Kelvin–Helmholtz shear instability (occurs when Richardson number \(\textrm{Ri}<0.25\) unless baroclinic instability is active)
MRW: mixed regular Wwves (steady or periodic baroclinic waves, mixed with the presence of short wavelength capillary-gravity waves)
MIW: mixed irregular waves (chaotic baroclinic waves, mixed with the presence of short wavelength capillary-gravity waves)

The movies below show examples of each of these regimes. 

Citations

Hart, J. E.; Kittelman, S. (1986). ‘A method for measuring interfacial wave fields in the laboratory’. Geophysical & Astrophysical Fluid Dynamics. 36 (2): 179–185. Bibcode:1986GApFD..36..179H. doi:10.1080/03091928608208802.
Lovegrove, A. F.; Moroz, I. M.; Read, P. L. (2001). ‘Bifurcations and instabilities in rotating two-layer fluids: I. \(f\)-plane’. Nonlinear Processes in Geophysics. 8: 21–36. Bibcode:2001NPGeo...8...21L. doi:10.5194/npg-8-21-2001.
Lovegrove, A. F.; Moroz, I. M.; Read, P. L. (2002). ‘Bifurcations and instabilities in rotating, two-layer fluids: II. \(\beta\)-plane’. Nonlinear Processes in Geophysics. 9 (3/4): 289–309. Bibcode:2002NPGeo...9..289L. doi:10.5194/npg-9-289-2002. S2CID 2000806.
Lovegrove, A. F.; Read, P. L.; Richards, C. J. (2000). ‘Generation of inertia-gravity waves in a baroclinically unstable fluid’. Quarterly Journal of the Royal Meteorological Society. 126 (570): 3233–3254. Bibcode:2000QJRMS.126.3233L. doi:10.1002/qj.49712657012. S2CID 122250300.
Phillips, N. A. (1951). ‘A simple three-dimensional model for the study of large-scale extratropical flow patterns’. Journal of Meteorology. 8 (6): 381. Bibcode:1951JAtS....8..381P. doi:10.1175/1520-0469(1951)008<0381:ASTDMF>2.0.CO;2. ISSN 1520-0469.
Williams, P. D.; Haine, T. W. N.; Read, P. L. (2005). ‘On the generation mechanisms of short-scale unbalanced modes in rotating two-layer flows with vertical shear’. Journal of Fluid Mechanics. 528: 1–22. Bibcode:2005JFM...528....1W. doi:10.1017/S0022112004002873. S2CID 53539087.
Williams, P. D.; Read, P. L.; Haine, T. W. N. (2004). ‘A calibrated, non-invasive method for measuring the internal interface height field at high resolution in the rotating, two-layer annulus’. Geophysical & Astrophysical Fluid Dynamics. 98 (6): 453–471. Bibcode:2004GApFD..98..453W. doi:10.1080/03091920412331296366. S2CID 120128486.