Baroclinic instabilities in the ocean

Last updated

A baroclinic instability is a fluid dynamical instability of fundamental importance in the atmosphere and ocean. It can lead to the formation of transient mesoscale eddies, with a horizontal scale of 10-100 km. [1] [2] In contrast, flows on the largest scale in the ocean are described as ocean currents, the largest scale eddies are mostly created by shearing of two ocean currents and static mesoscale eddies are formed by the flow around an obstacle (as seen in the animation on eddy (fluid dynamics). [2] [3] Mesoscale eddies are circular currents with swirling motion and account for approximately 90% of the ocean's total kinetic energy. [4] [5] Therefore, they are key in mixing and transport of for example heat, salt and nutrients.

Contents

In a baroclinic medium, the density depends on both the temperature and pressure. [6] The effect of the temperature on the density allows lines of equal density (isopycnals) and lines of equal pressure (isobars) to intersect. This is in contrast to a barotropic fluid, in which the density is only a function of pressure. For this barotropic case, isobars and isopycnals are parallel. The intersecting of isobars and isopycnals in a baroclinic medium may cause baroclinic instabilities to occur by the process of sloping convection. The sizes of baroclinic instabilities and therefore also the eddies they create scale with the Rossby radius of deformation, which strongly varies with latitude for the ocean.

Instability and eddy generation

A schematic on a baroclinic fluid with sloping isopycnals, intersecting with isobars on the Northern hemisphere, showing the process of sloping convection and formation of a baroclinic instability. When a fluid parcel is perturbed from its steady state location A to location B, it will be surrounded by a fluid with a lower density and the parcel will sink down to its original equilibrium location; the fluid parcel is now stable. However, when a parcel is displaced to location C, it is surrounded by fluid with a higher density than the parcel itself. The parcel will now float up even further, a small perturbation grows into a larger one and a baroclinic instability is formed. Baroclinic instability and sloping convection.jpg
A schematic on a baroclinic fluid with sloping isopycnals, intersecting with isobars on the Northern hemisphere, showing the process of sloping convection and formation of a baroclinic instability. When a fluid parcel is perturbed from its steady state location A to location B, it will be surrounded by a fluid with a lower density and the parcel will sink down to its original equilibrium location; the fluid parcel is now stable. However, when a parcel is displaced to location C, it is surrounded by fluid with a higher density than the parcel itself. The parcel will now float up even further, a small perturbation grows into a larger one and a baroclinic instability is formed.

In a baroclinic fluid, the thermal-wind balance holds, which is a combination of the geostrophic balance and the hydrostatic balance. This implies that isopycnals can slope with respect to the isobars. Furthermore, this also results in changing horizontal velocities with height as a result of horizontal temperature and therefore density gradients.

Under the thermal-wind balance, geostrophic balance and hydrostatic balance, a flow is in equilibrium. However, this is not the equilibrium of least energy. [7] A reduction in slope of the isopycnals would lower the center of gravity and therefore also the potential energy. It would also reduce the pressure gradient, leading to an increase in the kinetic energy. However, under the thermal-wind balance, a decrease in slope of the isopycnals cannot occur spontaneously. It requires a change of potential vorticity. [7] Under certain conditions, slight perturbations of the equilibrium under the thermal-wind balance may increase, leading to larger perturbations from the initial state and thus the growth of an instability.

It is often considered that baroclinic instability is the mechanism which extracts potential energy stored in horizontal density gradients and uses this "eddy potential energy" to drive eddies. [8]

Sloping convection

These baroclinic instabilities may be initiated by the process of 'sloping convection' or 'slanted thermal convection'. To understand this, consider a fluid in steady state and under the thermal-wind balance. Initially, a fluid parcel is at location A. The fluid parcel is slightly perturbed to location B, while still retaining its original density. Therefore, the fluid parcel is now in a location with a lower density than itself and the parcel will just sink down to its original position; the fluid parcel is now stable. However, when a parcel displaced to location C, it is surrounded by fluid with a higher density than the parcel itself. Due to its relatively low density with respect to its surroundings, the parcel will float up even further. Now a small perturbation grows into a larger one, which implies a baroclinic instability.

A criterion for an instability to occur can be defined. As stated before, in a baroclinic fluid, the thermal-wind balance holds, which implies the following two relations:

and ,

where is the density and , and are the spatial coordinates in the horizontal (latitudinal and longitudinal) and vertical direction, respectively. and represent the horizontal (zonal and meridional) components of the velocity vector in the - and -direction, respectively. Now thus and are the two horizontal density gradients. is the gravitational acceleration at the surface of the Earth and the Coriolis parameter.

Therefore a horizontal density gradient in the -direction leads to a gradient in horizontal flow velocity over depth .

The slope of the displacement is defined as

,

where and are the horizontal and vertical velocities of the perturbation, respectively.

An instability now occurs when the slope of the displacement is smaller than the slope of the isopycnals. The isopycnals can be mathematically described as . Now this results in an instability when:

From now on, only a two layer system with and the slopes of the top and bottom layer, respectively, is considered to simplify the problem. This is now similar to the classic Philips model. [9] From the thermal-wind balance it now follows that

where is the reduced gravity and the Coriolis-parameter at the equator according to the beta-plane approximation.

Performing a scale analysis on the slope of the perturbation allows to assign physical quantities to this mathematical problem. This now results in

,

where is the scale height, the horizontal length scale, and is the Rossby-parameter.

From this it can be stated that an instability occurs when

or ,

where is the reduced gravity and   is the velocity difference between the lower and upper layer. This criterion can be used to identify whether a small perturbation will grow into a larger one and thus whether an instability is expected to occur. From this it follows that you need some kind of shear to obtain an instability, it is easier to get an instability for long waves (perturbations) with large , and the and therefore the beta-effect is stabilizing. [7]

Furthermore, for the baroclinic Rossby radius of deformation it holds that . Now the instability criteria simplify to

or .

From this analysis it also follows that baroclinic instabilities are important for small Rossby numbers, where .

Observations of Baroclinic instabilities and eddies

Recently, many observations on mesoscale eddies in the ocean have been made using sea surface height data from altimeters. [10] It has been shown that regions with the highest growth rate of baroclinic instabilities indeed match the regions which are rich in eddies. [11] Furthermore, also the trajectories of both cyclonic and anticyclonic eddies can be studied. [10] From this it follows that there are approximately the same number of cyclonic and anticyclonic eddies observed and therefore it is concluded that the generation of these two types is very similar. However, when considering longer lived eddies, they found that anticyclonic eddies clearly dominate. This implies that cyclonic eddies are less stable and therefore decay more rapidly. [10] In addition, there are no eddies present above shallows in the ocean due to topographic steering as a result of the Taylor–Proudman theorem. Lastly, extremely long lived eddies with lifetimes over 1.5 to 2 years are only found in gyres, most likely because the background flow is small here.

Four different types of Baroclinic instabilities can be distinguished: [12]

These four types are based on classical models (the classic Eady Model, [13] the Charney model, and the Phillips model, [14] respectively), but can also be distinguished from observations. Overall, from the observed baroclinic instabilities 47% is the Charney surface type, 33% the Phillips type, 13% the Eady type and only 7% the Charney bottom type. [12] These different types of Baroclinic instabilities all lead to different types of eddies. Important here is ψ, which is the absolute value of the complex eigenfunction of the stream function of the horizontal velocity. [12] It represents the vertical structure of the Baroclinic instability and ranges from 0, which implies a very low chance of an instability of this type and thus also eddy to form, to 1, which means a high chance.

The Eady type has a maximum ψ of one at the top and bottom, and a minimum around 0.5 halfway the total depth. For this type of model, an eddy thus occurs at both the surface and bottom of the ocean. It is therefore also called the surface- and bottom-intensified type and found mainly at high latitudes. [12] The Charney surface type is surface-intensified and has a maximum ψ at the surface, whereas the Charney bottom type only shows baroclinic instabilities at the bottom. For the Charney bottom type ψ is also at the surface and increases to one over increasing depth. The Charney surface type is found in the subtropics, whereas the Charney bottom type is present at high latitudes. Lastly, for the Phillips type, ψ is zero at the surface, strongly increases to one just below the surface, and then slowly decreases again to zero for increasing depths. The location of these Phillips type instabilities agree with the occurrence of subsurface eddies, again supporting the idea that the Baroclinic instabilities lead to the formation of eddies. [12] They are mostly found in the tropics and the eastern return flow of the subtropical gyres.

It was found that the type of Baroclinic instability present also depends on the mean background flow. [12] An Eady type is preferred for a strong eastward mean flow in the upper ocean, and a weak westward flow in the deeper ocean. For the Charney bottom type this is similar, but now the westward flow in the deeper ocean is found to be stronger. The Charney surface and Phillips types exist for weaker background flows, also explaining why these are dominant in the ocean gyres.

Related Research Articles

<span class="mw-page-title-main">Navier–Stokes equations</span> Equations describing the motion of viscous fluid substances

The Navier–Stokes equations are partial differential equations which describe the motion of viscous fluid substances. They were named after French engineer and physicist Claude-Louis Navier and the Irish physicist and mathematician George Gabriel Stokes. They were developed over several decades of progressively building the theories, from 1822 (Navier) to 1842–1850 (Stokes).

In fluid mechanics, the Grashof number is a dimensionless number which approximates the ratio of the buoyancy to viscous forces acting on a fluid. It frequently arises in the study of situations involving natural convection and is analogous to the Reynolds number.

<span class="mw-page-title-main">Potential flow</span> Velocity field as the gradient of a scalar function

In fluid dynamics, potential flow is the ideal flow pattern of an inviscid fluid. Potential flows are described and determined by mathematical methods.

<span class="mw-page-title-main">Baroclinity</span> Measure of misalignment between the gradients of pressure and density in a fluid

In fluid dynamics, the baroclinity of a stratified fluid is a measure of how misaligned the gradient of pressure is from the gradient of density in a fluid. In meteorology a baroclinic flow is one in which the density depends on both temperature and pressure. A simpler case, barotropic flow, allows for density dependence only on pressure, so that the curl of the pressure-gradient force vanishes.

<span class="mw-page-title-main">Euler equations (fluid dynamics)</span> Set of quasilinear hyperbolic equations governing adiabatic and inviscid flow

In fluid dynamics, the Euler equations are a set of quasilinear partial differential equations governing adiabatic and inviscid flow. They are named after Leonhard Euler. In particular, they correspond to the Navier–Stokes equations with zero viscosity and zero thermal conductivity.

<span class="mw-page-title-main">Hydrostatics</span> Branch of fluid mechanics that studies fluids at rest

Fluid statics or hydrostatics is the branch of fluid mechanics that studies fluids at hydrostatic equilibrium and "the pressure in a fluid or exerted by a fluid on an immersed body".

<span class="mw-page-title-main">Ekman spiral</span> Velocity profile of wind driven current with depth

The Ekman spiral is an arrangement of ocean currents: the directions of horizontal current appear to twist as the depth changes. The oceanic wind driven Ekman spiral is the result of a force balance created by a shear stress force, Coriolis force and the water drag. This force balance gives a resulting current of the water different from the winds. In the ocean, there are two places where the Ekman spiral can be observed. At the surface of the ocean, the shear stress force corresponds with the wind stress force. At the bottom of the ocean, the shear stress force is created by friction with the ocean floor. This phenomenon was first observed at the surface by the Norwegian oceanographer Fridtjof Nansen during his Fram expedition. He noticed that icebergs did not drift in the same direction as the wind. His student, the Swedish oceanographer Vagn Walfrid Ekman, was the first person to physically explain this process.

<span class="mw-page-title-main">Eddy (fluid dynamics)</span> Swirling of a fluid and the reverse current created when the fluid is in a turbulent flow regime

In fluid dynamics, an eddy is the swirling of a fluid and the reverse current created when the fluid is in a turbulent flow regime. The moving fluid creates a space devoid of downstream-flowing fluid on the downstream side of the object. Fluid behind the obstacle flows into the void creating a swirl of fluid on each edge of the obstacle, followed by a short reverse flow of fluid behind the obstacle flowing upstream, toward the back of the obstacle. This phenomenon is naturally observed behind large emergent rocks in swift-flowing rivers.

<span class="mw-page-title-main">Ekman layer</span> Force equilibrium layer in a liquid

The Ekman layer is the layer in a fluid where there is a force balance between pressure gradient force, Coriolis force and turbulent drag. It was first described by Vagn Walfrid Ekman. Ekman layers occur both in the atmosphere and in the ocean.

In fluid mechanics, potential vorticity (PV) is a quantity which is proportional to the dot product of vorticity and stratification. This quantity, following a parcel of air or water, can only be changed by diabatic or frictional processes. It is a useful concept for understanding the generation of vorticity in cyclogenesis, especially along the polar front, and in analyzing flow in the ocean.

<span class="mw-page-title-main">Eddy diffusion</span> Mixing of fluids due to eddy currents

In fluid dynamics, eddy diffusion, eddy dispersion, or turbulent diffusion is a process by which fluid substances mix together due to eddy motion. These eddies can vary widely in size, from subtropical ocean gyres down to the small Kolmogorov microscales, and occur as a result of turbulence. The theory of eddy diffusion was first developed by Sir Geoffrey Ingram Taylor.

<span class="mw-page-title-main">Shallow water equations</span> Set of partial differential equations that describe the flow below a pressure surface in a fluid

The shallow-water equations (SWE) are a set of hyperbolic partial differential equations that describe the flow below a pressure surface in a fluid. The shallow-water equations in unidirectional form are also called Saint-Venant equations, after Adhémar Jean Claude Barré de Saint-Venant.

In physical oceanography and fluid dynamics, the wind stress is the shear stress exerted by the wind on the surface of large bodies of water – such as oceans, seas, estuaries and lakes. When wind is blowing over a water surface, the wind applies a wind force on the water surface. The wind stress is the component of this wind force that is parallel to the surface per unit area. Also, the wind stress can be described as the flux of horizontal momentum applied by the wind on the water surface. The wind stress causes a deformation of the water body whereby wind waves are generated. Also, the wind stress drives ocean currents and is therefore an important driver of the large-scale ocean circulation. The wind stress is affected by the wind speed, the shape of the wind waves and the atmospheric stratification. It is one of the components of the air–sea interaction, with others being the atmospheric pressure on the water surface, as well as the exchange of energy and mass between the water and the atmosphere.

In fluid dynamics, Luke's variational principle is a Lagrangian variational description of the motion of surface waves on a fluid with a free surface, under the action of gravity. This principle is named after J.C. Luke, who published it in 1967. This variational principle is for incompressible and inviscid potential flows, and is used to derive approximate wave models like the mild-slope equation, or using the averaged Lagrangian approach for wave propagation in inhomogeneous media.

In fluid dynamics, Airy wave theory gives a linearised description of the propagation of gravity waves on the surface of a homogeneous fluid layer. The theory assumes that the fluid layer has a uniform mean depth, and that the fluid flow is inviscid, incompressible and irrotational. This theory was first published, in correct form, by George Biddell Airy in the 19th century.

Ocean dynamics define and describe the flow of water within the oceans. Ocean temperature and motion fields can be separated into three distinct layers: mixed (surface) layer, upper ocean, and deep ocean.

<span class="mw-page-title-main">Radiation stress</span> Term in physical oceanography

In fluid dynamics, the radiation stress is the depth-integrated – and thereafter phase-averaged – excess momentum flux caused by the presence of the surface gravity waves, which is exerted on the mean flow. The radiation stresses behave as a second-order tensor.

<span class="mw-page-title-main">Rayleigh–Kuo criterion</span> Stability condition for fluids

The Rayleigh–Kuo criterion is a stability condition for a fluid. This criterion determines whether or not a barotropic instability can occur, leading to the presence of vortices. The Kuo criterion states that for barotropic instability to occur, the gradient of the absolute vorticity must change its sign at some point within the boundaries of the current. Note that this criterion is a necessary condition, so if it does not hold it is not possible for a barotropic instability to form. But it is not a sufficient condition, meaning that if the criterion is met, this does not automatically mean that the fluid is unstable. If the criterion is not met, it is certain that the flow is stable.

Open ocean convection is a process in which the mesoscale ocean circulation and large, strong winds mix layers of water at different depths. Fresher water lying over the saltier or warmer over the colder leads to the stratification of water, or its separation into layers. Strong winds cause evaporation, so the ocean surface cools, weakening the stratification. As a result, the surface waters are overturned and sink while the "warmer" waters rise to the surface, starting the process of convection. This process has a crucial role in the formation of both bottom and intermediate water and in the large-scale thermohaline circulation, which largely determines global climate. It is also an important phenomena that controls the intensity of the Atlantic Meridional Overturning Circulation (AMOC).

Eddy pumping is a component of mesoscale eddy-induced vertical motion in the ocean. It is a physical mechanism through which vertical motion is created from variations in an eddy's rotational strength. Cyclonic (Anticyclonic) eddies lead primarily to upwelling (downwelling) in the Northern Hemisphere and vice versa in the Southern hemisphere. It is a key mechanism driving biological and biogeochemical processes in the ocean such as algal blooms and the carbon cycle.

References

  1. GFDL NOAA. "Ocean Mesoscale Eddies" . Retrieved 5 June 2021.
  2. 1 2 George, Tom M.; Manucharyan, Georgy E.; Thompson, Andrew F. (2021-02-05). "Deep learning to infer eddy heat fluxes from sea surface height patterns of mesoscale turbulence". Nature Communications. 12 (1): 800. Bibcode:2021NatCo..12..800G. doi:10.1038/s41467-020-20779-9. ISSN   2041-1723. PMC   7865057 . PMID   33547299.
  3. NOAA (26 February 2021). "What is an eddy?" . Retrieved 5 June 2021.
  4. Wunsch, Carl; Ferrari, Raffaele (January 2004). "Vertical Mixing, Energy, and the General Circulation of the Oceans". Annual Review of Fluid Mechanics. 36 (1): 281–314. Bibcode:2004AnRFM..36..281W. doi:10.1146/annurev.fluid.36.050802.122121. ISSN   0066-4189.
  5. Venaille, Antoine; Vallis, Geoffrey K.; Smith, K. Shafer (2011-09-01). "Baroclinic Turbulence in the Ocean: Analysis with Primitive Equation and Quasigeostrophic Simulations". Journal of Physical Oceanography. 41 (9): 1605–1623. Bibcode:2011JPO....41.1605V. doi: 10.1175/jpo-d-10-05021.1 . ISSN   0022-3670.
  6. "International geophysics series", An Introduction to Dynamic Meteorology, International Geophysics, vol. 88, Elsevier, 2004, pp. 531–535, doi:10.1016/s0074-6142(04)80057-3, ISBN   978-0-12-354015-7 , retrieved 2021-05-17
  7. 1 2 3 Cushman-Roisin, Benoit; Beckers, Jean-Marie (2011), "Equations Governing Geophysical Flows", International Geophysics, Elsevier, pp. 99–129, doi:10.1016/b978-0-12-088759-0.00004-3, ISBN   978-0-12-088759-0 , retrieved 2021-06-19
  8. Gill, A.E.; Green, J.S.A.; Simmons, A.J. (July 1974). "Energy partition in the large-scale ocean circulation and the production of mid-ocean eddies". Deep Sea Research and Oceanographic Abstracts. 21 (7): 499–528. Bibcode:1974DSRA...21..499G. doi:10.1016/0011-7471(74)90010-2. ISSN   0011-7471.
  9. Phillips, Norman A. (January 1954). "Energy Transformations and Meridional Circulations associated with simple Baroclinic Waves in a two-level, Quasi-geostrophic Model". Tellus. 6 (3): 274–286. Bibcode:1954Tell....6..274P. doi: 10.3402/tellusa.v6i3.8734 . ISSN   0040-2826.
  10. 1 2 3 Chelton, Dudley B.; Schlax, Michael G.; Samelson, Roger M. (October 2011). "Global observations of nonlinear mesoscale eddies". Progress in Oceanography. 91 (2): 167–216. Bibcode:2011PrOce..91..167C. doi:10.1016/j.pocean.2011.01.002. ISSN   0079-6611.
  11. Smith, K. Shafer (2007-09-01). "The geography of linear baroclinic instability in Earth's oceans". Journal of Marine Research. 65 (5): 655–683. doi:10.1357/002224007783649484. ISSN   0022-2402.
  12. 1 2 3 4 5 6 Feng, Ling; Liu, Chuanyu; Köhl, Armin; Stammer, Detlef; Wang, Fan (March 2021). "Four Types of Baroclinic Instability Waves in the Global Oceans and the Implications for the Vertical Structure of Mesoscale Eddies". Journal of Geophysical Research: Oceans. 126 (3). Bibcode:2021JGRC..12616966F. doi: 10.1029/2020jc016966 . ISSN   2169-9275.
  13. Eady, E. T. (August 1949). "Long Waves and Cyclone Waves". Tellus. 1 (3): 33–52. Bibcode:1949Tell....1c..33E. doi:10.1111/j.2153-3490.1949.tb01265.x. ISSN   0040-2826.
  14. Phillips, Norman A. (January 1954). "Energy Transformations and Meridional Circulations associated with simple Baroclinic Waves in a two-level, Quasi-geostrophic Model". Tellus. 6 (3): 274–286. Bibcode:1954Tell....6..274P. doi: 10.3402/tellusa.v6i3.8734 . ISSN   0040-2826.