Rossby wave

Last updated

Meanders of the Northern Hemisphere's jet stream developing around the northern polar vortex (a, b) and finally detaching a "drop" of cold air (c). Orange: warmer masses of air; pink: jet stream; blue: colder masses of air. Jetstream - Rossby Waves - N hemisphere.svg
Meanders of the Northern Hemisphere's jet stream developing around the northern polar vortex (a, b) and finally detaching a "drop" of cold air (c). Orange: warmer masses of air; pink: jet stream; blue: colder masses of air.

Rossby waves, also known as planetary waves, are a type of inertial wave naturally occurring in rotating fluids. [1] They were first identified by Sweden-born American meteorologist Carl-Gustaf Arvid Rossby in the Earth's atmosphere in 1939. They are observed in the atmospheres and oceans of Earth and other planets, owing to the rotation of Earth or of the planet involved. Atmospheric Rossby waves on Earth are giant meanders in high-altitude winds that have a major influence on weather. These waves are associated with pressure systems and the jet stream (especially around the polar vortices). [2] Oceanic Rossby waves move along the thermocline: the boundary between the warm upper layer and the cold deeper part of the ocean.

Contents

Rossby wave types

Atmospheric waves

Sketches of Rossby waves' fundamental principles. a and b The restoring force. c-e The waveform's velocity. In a, an air parcel follows along latitude
ph
0
{\displaystyle \varphi _{0}}
at an eastward velocity
v
E
{\displaystyle v_{E}}
with a meridional acceleration
a
N
=
0
{\displaystyle a_{N}=0}
when the pressure gradient force balances the Coriolis force. In b, when the parcel encounters a small displacement
d
ph
{\displaystyle \delta \varphi }
in latitude, the Coriolis force's gradient imposes a meridional acceleration
a
N
{\displaystyle a_{N}}
that always points against
d
ph
{\displaystyle \delta \varphi }
when
v
E
>
0
{\displaystyle v_{E}>0}
. Here,
O
{\displaystyle \Omega }
denotes the Earth's angular frequency and
a
N
{\displaystyle a_{N}}
is the northward Coriolis acceleration. While the parcel meanders along the blue arrowed line
l
{\displaystyle l}
in b , its waveform travels westward as sketched in c. The absolute vorticity composes the planetary vorticity
f
{\displaystyle f}
and the relative vorticity
z
{\displaystyle \zeta }
, reflecting the Earth's rotation and the parcel's rotation with respect to the Earth, respectively. The conservation of absolute vorticity
e
{\displaystyle \eta }
determines a southward gradient of
z
{\displaystyle \zeta }
, as denoted by the red shadow in c. The gradient's projection along the flow path
l
{\displaystyle l}
is typically not zero and would cause a tangential velocity
v
t
{\displaystyle v_{t}}
. As an example, the path
l
{\displaystyle l}
in c is zoomed in at two green crosses, displayed in d and e. These two crosses are associated with positive and negative gradients of
z
{\displaystyle \zeta }
along
l
{\displaystyle l}
, respectively, as denoted by the red and pink arrows in d and e. The black arrows
v
t
{\displaystyle v_{t}}
denote the vector sums of the red and pink arrows bordering the crosses, both of which project zonally westward. The parcels at these crosses drift toward the green points in c and, visually, the path
l
{\displaystyle l}
drifts westward toward the dotted line. Sketches of Rossby wave's fundamental principles..png
Sketches of Rossby waves’ fundamental principles. a and b The restoring force. ce The waveform’s velocity. In a, an air parcel follows along latitude  at an eastward velocity  with a meridional acceleration  when the pressure gradient force balances the Coriolis force. In b, when the parcel encounters a small displacement  in latitude, the Coriolis force’s gradient imposes a meridional acceleration  that always points against  when . Here,  denotes the Earth’s angular frequency and  is the northward Coriolis acceleration. While the parcel meanders along the blue arrowed line  in b , its waveform travels westward as sketched in c. The absolute vorticity composes the planetary vorticity  and the relative vorticity , reflecting the Earth’s rotation and the parcel’s rotation with respect to the Earth, respectively. The conservation of absolute vorticity  determines a southward gradient of , as denoted by the red shadow in c. The gradient’s projection along the flow path  is typically not zero and would cause a tangential velocity . As an example, the path  in c is zoomed in at two green crosses, displayed in d and e. These two crosses are associated with positive and negative gradients of  along , respectively, as denoted by the red and pink arrows in d and e. The black arrows  denote the vector sums of the red and pink arrows bordering the crosses, both of which project zonally westward. The parcels at these crosses drift toward the green points in c and, visually, the path  drifts westward toward the dotted line.

Atmospheric Rossby waves result from the conservation of potential vorticity and are influenced by the Coriolis force and pressure gradient. [3] The image on the left sketches fundamental principles of the wave, e.g., its restoring force and westward phase velocity. The rotation causes fluids to turn to the right as they move in the northern hemisphere and to the left in the southern hemisphere. For example, a fluid that moves from the equator toward the north pole will deviate toward the east; a fluid moving toward the equator from the north will deviate toward the west. These deviations are caused by the Coriolis force and conservation of potential vorticity which leads to changes of relative vorticity. This is analogous to conservation of angular momentum in mechanics. In planetary atmospheres, including Earth, Rossby waves are due to the variation in the Coriolis effect with latitude.

One can identify a terrestrial Rossby wave as its phase velocity, marked by its wave crest, always has a westward component.[ citation needed ] However, the collected set of Rossby waves may appear to move in either direction with what is known as its group velocity. In general, shorter waves have an eastward group velocity and long waves a westward group velocity.

The terms "barotropic" and "baroclinic" are used to distinguish the vertical structure of Rossby waves. Barotropic Rossby waves do not vary in the vertical[ clarification needed ], and have the fastest propagation speeds. The baroclinic wave modes, on the other hand, do vary in the vertical. They are also slower, with speeds of only a few centimeters per second or less. [4]

Most investigations of Rossby waves have been done on those in Earth's atmosphere. Rossby waves in the Earth's atmosphere are easy to observe as (usually 4–6) large-scale meanders of the jet stream. When these deviations become very pronounced, masses of cold or warm air detach, and become low-strength cyclones and anticyclones, respectively, and are responsible for day-to-day weather patterns at mid-latitudes. The action of Rossby waves partially explains why eastern continental edges in the Northern Hemisphere, such as the Northeast United States and Eastern Canada, are colder than Western Europe at the same latitudes, [5] and why the Mediterranean is dry during summer (Rodwell–Hoskins mechanism). [6]

Poleward-propagating atmospheric waves

Deep convection (heat transfer) to the troposphere is enhanced over very warm sea surfaces in the tropics, such as during El Niño events. This tropical forcing generates atmospheric Rossby waves that have a poleward and eastward migration.

Poleward-propagating Rossby waves explain many of the observed statistical connections between low- and high-latitude climates. [7] One such phenomenon is sudden stratospheric warming. Poleward-propagating Rossby waves are an important and unambiguous part of the variability in the Northern Hemisphere, as expressed in the Pacific North America pattern. Similar mechanisms apply in the Southern Hemisphere and partly explain the strong variability in the Amundsen Sea region of Antarctica. [8] In 2011, a Nature Geoscience study using general circulation models linked Pacific Rossby waves generated by increasing central tropical Pacific temperatures to warming of the Amundsen Sea region, leading to winter and spring continental warming of Ellsworth Land and Marie Byrd Land in West Antarctica via an increase in advection. [9]

Rossby waves on other planets

Atmospheric Rossby waves, like Kelvin waves, can occur on any rotating planet with an atmosphere. The Y-shaped cloud feature on Venus is attributed to Kelvin and Rossby waves. [10]

Oceanic waves

Oceanic Rossby waves are large-scale waves within an ocean basin. They have a low amplitude, in the order of centimetres (at the surface) to metres (at the thermocline), compared with atmospheric Rossby waves which are in the order of hundreds of kilometres. They may take months to cross an ocean basin. They gain momentum from wind stress at the ocean surface layer and are thought to communicate climatic changes due to variability in forcing, due to both the wind and buoyancy. Off-equatorial Rossby waves are believed to propagate through eastward-propagating Kelvin waves that upwell against Eastern Boundary Currents, while equatorial Kelvin waves are believed to derive some of their energy from the reflection of Rossby waves against Western Boundary Currents. [11]

Both barotropic and baroclinic waves cause variations of the sea surface height, although the length of the waves made them difficult to detect until the advent of satellite altimetry. Satellite observations have confirmed the existence of oceanic Rossby waves. [12]

Baroclinic waves also generate significant displacements of the oceanic thermocline, often of tens of meters. Satellite observations have revealed the stately progression of Rossby waves across all the ocean basins, particularly at low- and mid-latitudes. Due to the beta effect, transit times of Rossby waves increase with latitude. In a basin like the Pacific, waves travelling at the equator may take months, while closer to the poles transit may take decades. [13]

Rossby waves have been suggested as an important mechanism to account for the heating of the ocean on Europa, a moon of Jupiter. [14]

Waves in astrophysical discs

Rossby wave instabilities are also thought to be found in astrophysical discs, for example, around newly forming stars. [15] [16]

Amplification of Rossby waves

It has been proposed that a number of regional weather extremes in the Northern Hemisphere associated with blocked atmospheric circulation patterns may have been caused by quasiresonant amplification of Rossby waves. Examples include the 2013 European floods, the 2012 China floods, the 2010 Russian heat wave, the 2010 Pakistan floods and the 2003 European heat wave. Even taking global warming into account, the 2003 heat wave would have been highly unlikely without such a mechanism.

Normally freely travelling synoptic-scale Rossby waves and quasistationary planetary-scale Rossby waves exist in the mid-latitudes with only weak interactions. The hypothesis, proposed by Vladimir Petoukhov, Stefan Rahmstorf, Stefan Petri, and Hans Joachim Schellnhuber, is that under some circumstances these waves interact to produce the static pattern. For this to happen, they suggest, the zonal (east-west) wave number of both types of wave should be in the range 6–8, the synoptic waves should be arrested within the troposphere (so that energy does not escape to the stratosphere) and mid-latitude waveguides should trap the quasistationary components of the synoptic waves. In this case the planetary-scale waves may respond unusually strongly to orography and thermal sources and sinks because of "quasiresonance". [17]

A 2017 study by Mann, Rahmstorf, et al. connected the phenomenon of anthropogenic Arctic amplification to planetary wave resonance and extreme weather events. [18]

Mathematical definitions

Free barotropic Rossby waves under a zonal flow with linearized vorticity equation

To start with, a zonal mean flow, U, can be considered to be perturbed where U is constant in time and space. Let be the total horizontal wind field, where u and v are the components of the wind in the x- and y- directions, respectively. The total wind field can be written as a mean flow, U, with a small superimposed perturbation, u and v.

The perturbation is assumed to be much smaller than the mean zonal flow.

The relative vorticity and the perturbations and can be written in terms of the stream function (assuming non-divergent flow, for which the stream function completely describes the flow):

Considering a parcel of air that has no relative vorticity before perturbation (uniform U has no vorticity) but with planetary vorticity f as a function of the latitude, perturbation will lead to a slight change of latitude, so the perturbed relative vorticity must change in order to conserve potential vorticity. Also the above approximation U >> u' ensures that the perturbation flow does not advect relative vorticity.

with . Plug in the definition of stream function to obtain:

Using the method of undetermined coefficients one can consider a traveling wave solution with zonal and meridional wavenumbers k and , respectively, and frequency :

This yields the dispersion relation:

The zonal (x-direction) phase speed and group velocity of the Rossby wave are then given by

where c is the phase speed, cg is the group speed, U is the mean westerly flow, is the Rossby parameter, k is the zonal wavenumber, and is the meridional wavenumber. It is noted that the zonal phase speed of Rossby waves is always westward (traveling east to west) relative to mean flow U, but the zonal group speed of Rossby waves can be eastward or westward depending on wavenumber.

Rossby parameter

The Rossby parameter is defined as the rate of change of the Coriolis frequency along the meridional direction:

where is the latitude, ω is the angular speed of the Earth's rotation, and a is the mean radius of the Earth.

If , there will be no Rossby waves; Rossby waves owe their origin to the gradient of the tangential speed of the planetary rotation (planetary vorticity). A "cylinder" planet has no Rossby waves. It also means that at the equator of any rotating, sphere-like planet, including Earth, one will still have Rossby waves, despite the fact that , because . These are known as Equatorial Rossby waves.

See also

Related Research Articles

In particle physics, the Dirac equation is a relativistic wave equation derived by British physicist Paul Dirac in 1928. In its free form, or including electromagnetic interactions, it describes all spin-12 massive particles, called "Dirac particles", such as electrons and quarks for which parity is a symmetry. It is consistent with both the principles of quantum mechanics and the theory of special relativity, and was the first theory to account fully for special relativity in the context of quantum mechanics. It was validated by accounting for the fine structure of the hydrogen spectrum in a completely rigorous way.

<span class="mw-page-title-main">Gravity wave</span> Wave in or at the interface between fluids where gravity is the main equilibrium force

In fluid dynamics, gravity waves are waves generated in a fluid medium or at the interface between two media when the force of gravity or buoyancy tries to restore equilibrium. An example of such an interface is that between the atmosphere and the ocean, which gives rise to wind waves.

<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.

A Kelvin wave is a wave in the ocean, a large lake or the atmosphere that balances the Earth's Coriolis force against a topographic boundary such as a coastline, or a waveguide such as the equator. A feature of a Kelvin wave is that it is non-dispersive, i.e., the phase speed of the wave crests is equal to the group speed of the wave energy for all frequencies. This means that it retains its shape as it moves in the alongshore direction over time.

<span class="mw-page-title-main">LSZ reduction formula</span> Connection between correlation functions and the S-matrix

In quantum field theory, the Lehmann–Symanzik–Zimmermann (LSZ) reduction formula is a method to calculate S-matrix elements from the time-ordered correlation functions of a quantum field theory. It is a step of the path that starts from the Lagrangian of some quantum field theory and leads to prediction of measurable quantities. It is named after the three German physicists Harry Lehmann, Kurt Symanzik and Wolfhart Zimmermann.

<span class="mw-page-title-main">Rayleigh–Taylor instability</span> Unstable behavior of two contacting fluids of different densities

The Rayleigh–Taylor instability, or RT instability, is an instability of an interface between two fluids of different densities which occurs when the lighter fluid is pushing the heavier fluid. Examples include the behavior of water suspended above oil in the gravity of Earth, mushroom clouds like those from volcanic eruptions and atmospheric nuclear explosions, supernova explosions in which expanding core gas is accelerated into denser shell gas, instabilities in plasma fusion reactors and inertial confinement fusion.

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">Boundary current</span> Ocean current with dynamics determined by the presence of a coastline

Boundary currents are ocean currents with dynamics determined by the presence of a coastline, and fall into two distinct categories: western boundary currents and eastern boundary currents.

Equatorial waves are oceanic and atmospheric waves trapped close to the equator, meaning that they decay rapidly away from the equator, but can propagate in the longitudinal and vertical directions. Wave trapping is the result of the Earth's rotation and its spherical shape which combine to cause the magnitude of the Coriolis force to increase rapidly away from the equator. Equatorial waves are present in both the tropical atmosphere and ocean and play an important role in the evolution of many climate phenomena such as El Niño. Many physical processes may excite equatorial waves including, in the case of the atmosphere, diabatic heat release associated with cloud formation, and in the case of the ocean, anomalous changes in the strength or direction of the trade winds.

Rossby-gravity waves are equatorially trapped waves, meaning that they rapidly decay as their distance increases away from the equator. These waves have the same trapping scale as Kelvin waves, more commonly known as the equatorial Rossby deformation radius. They always carry energy eastward, but their 'crests' and 'troughs' may propagate westward if their periods are long enough.

Equatorial Rossby waves, often called planetary waves, are very long, low frequency water waves found near the equator and are derived using the equatorial beta plane approximation.

The Eady model is an atmospheric model for baroclinic instability first posed by British meteorologist Eric Eady in 1949 based on his PhD work at Imperial College London.

Q-vectors are used in atmospheric dynamics to understand physical processes such as vertical motion and frontogenesis. Q-vectors are not physical quantities that can be measured in the atmosphere but are derived from the quasi-geostrophic equations and can be used in the previous diagnostic situations. On meteorological charts, Q-vectors point toward upward motion and away from downward motion. Q-vectors are an alternative to the omega equation for diagnosing vertical motion in the quasi-geostrophic equations.

The soler model is a quantum field theory model of Dirac fermions interacting via four fermion interactions in 3 spatial and 1 time dimension. It was introduced in 1938 by Dmitri Ivanenko and re-introduced and investigated in 1970 by Mario Soler as a toy model of self-interacting electron.

<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.

The Atlantic meridional overturning circulation (AMOC) is a large system of ocean currents, like a conveyor belt. It is driven by differences in temperature and salt content and it is an important component of the climate system. However, the AMOC is not a static feature of global circulation. It is sensitive to changes in temperature, salinity and atmospheric forcings. Climate reconstructions from δ18O proxies from Greenland reveal an abrupt transition in global temperature about every 1470 years. These changes may be due to changes in ocean circulation, which suggests that there are two equilibria possible in the AMOC. Stommel made a two-box model in 1961 which showed two different states of the AMOC are possible on a single hemisphere. Stommel’s result with an ocean box model has initiated studies using three dimensional ocean circulation models, confirming the existence of multiple equilibria in the AMOC.

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. 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. Mesoscale eddies are circular currents with swirling motion and account for approximately 90% of the ocean's total kinetic energy. Therefore, they are key in mixing and transport of for example heat, salt and nutrients.

In fluid mechanics, topographic steering is the effect of potential vorticity conservation on the motion of a fluid parcel. This means that the fluid parcels will not only react to physical obstacles in their path, but also to changes in topography or latitude. The two types of 'fluids' where topographic steering is mainly observed in daily life are air and water in respectively the atmosphere and the oceans. Examples of topographic steering can be found in, among other things, paths of low pressure systems and oceanic currents.

<span class="mw-page-title-main">Topographic Rossby waves</span> Description of waves in the ocean and atmosphere created by bottom irregularities

Topographic Rossby waves are geophysical waves that form due to bottom irregularities. For ocean dynamics, the bottom irregularities are on the ocean floor such as the mid-ocean ridge. For atmospheric dynamics, the other primary branch of geophysical fluid dynamics, the bottom irregularities are found on land, for example in the form of mountains. Topographic Rossby waves are one of two types of geophysical waves named after the meteorologist Carl-Gustaf Rossby. The other type of Rossby waves are called planetary Rossby waves and have a different physical origin. Planetary Rossby waves form due to the changing Coriolis parameter over the earth. Rossby waves are quasi-geostrophic, dispersive waves. This means that not only the Coriolis force and the pressure-gradient force influence the flow, as in geostrophic flow, but also inertia.

The recharge oscillator model for El Niño–Southern Oscillation (ENSO) is a theory described for the first time in 1997 by Jin., which explains the periodical variation of the sea surface temperature (SST) and thermocline depth that occurs in the central equatorial Pacific Ocean. The physical mechanisms at the basis of this oscillation are periodical recharges and discharges of the zonal mean equatorial heat content, due to ocean-atmosphere interaction. Other theories have been proposed to model ENSO, such as the delayed oscillator, the western Pacific oscillator and the advective reflective oscillator. A unified and consistent model has been proposed by Wang in 2001, in which the recharge oscillator model is included as a particular case.

References

  1. "What is a Rossby wave?". National Oceanic and Atmospheric Administration.
  2. Holton, James R. (2004). Dynamic Meteorology. Elsevier. p. 347. ISBN   978-0-12-354015-7.
  3. 1 2 He, Maosheng; Forbes, Jeffrey M. (7 December 2022). "Rossby wave second harmonic generation observed in the middle atmosphere". Nature Communications. 13 (1): 7544. Bibcode:2022NatCo..13.7544H. doi:10.1038/s41467-022-35142-3. ISSN   2041-1723. PMC   9729661 . PMID   36476614. Creative Commons by small.svg  This article incorporates textfrom this source, which is available under the CC BY 4.0 license.
  4. Shepherd, Theodore G. (October 1987). "Rossby waves and two-dimensional turbulence in a large-scale zonal jet". Journal of Fluid Mechanics. 183: 467–509. Bibcode:1987JFM...183..467S. doi:10.1017/S0022112087002738. S2CID   9289503.
  5. Kaspi, Yohai; Schneider, Tapio (March 2011). "Winter cold of eastern continental boundaries induced by warm ocean waters" (PDF). Nature. 471 (7340): 621–624. Bibcode:2011Natur.471..621K. doi:10.1038/nature09924. PMID   21455177. S2CID   4388818.
  6. Rodwell, Mark J.; Hoskins, Brian J. (1996). "Monsoons and the dynamics of deserts". Quarterly Journal of the Royal Meteorological Society. 122 (534): 1385–1404. Bibcode:1996QJRMS.122.1385R. doi:10.1002/qj.49712253408. ISSN   1477-870X.
  7. Hoskins, Brian J.; Karoly, David J. (June 1981). "The Steady Linear Response of a Spherical Atmosphere to Thermal and Orographic Forcing". Journal of the Atmospheric Sciences. 38 (6): 1179–1196. Bibcode:1981JAtS...38.1179H. doi: 10.1175/1520-0469(1981)038<1179:TSLROA>2.0.CO;2 .
  8. Lachlan-Cope, Tom; Connolley, William (16 December 2006). "Teleconnections between the tropical Pacific and the Amundsen-Bellinghausens Sea: Role of the El Niño/Southern Oscillation". Journal of Geophysical Research: Atmospheres. 111 (D23). Bibcode:2006JGRD..11123101L. doi: 10.1029/2005JD006386 .
  9. Ding, Qinghua; Steig, Eric J.; Battisti, David S.; Küttel, Marcel (June 2011). "Winter warming in West Antarctica caused by central tropical Pacific warming". Nature Geoscience. 4 (6): 398–403. Bibcode:2011NatGe...4..398D. CiteSeerX   10.1.1.459.8689 . doi:10.1038/ngeo1129.
  10. Covey, Curt; Schubert, Gerald (November 1982). "Planetary-Scale Waves in the Venus Atmosphere". Journal of the Atmospheric Sciences. 39 (11): 2397–2413. Bibcode:1982JAtS...39.2397C. doi: 10.1175/1520-0469(1982)039<2397:PSWITV>2.0.CO;2 .
  11. Battisti, David S. (April 1989). "On the Role of Off-Equatorial Oceanic Rossby Waves during ENSO". Journal of Physical Oceanography. 19.4: 551–560.
  12. Chelton, D. B.; Schlax, M. G. (1996). "Global Observations of Oceanic Rossby Waves". Science. 272 (5259): 234. Bibcode:1996Sci...272..234C. doi:10.1126/science.272.5259.234. S2CID   126953559.
  13. Chelton, Dudley B.; Schlax, Michael B. (1996). "Global Observations of Oceanic Rossby Waves" (PDF). Science. 272 (5259): 234–238.
  14. Tyler, Robert H. (2008). "Strong ocean tidal flow and heating on moons of the outer planets". Nature. 456 (7223): 770–2. Bibcode:2008Natur.456..770T. doi:10.1038/nature07571. PMID   19079055. S2CID   205215528.
  15. Lovelace, R. V. E.; Li, H.; Colgate, S. A.; Nelson, A. F. (10 March 1999). "Rossby Wave Instability of Keplerian Accretion Disks". The Astrophysical Journal. 513 (2): 805–810. arXiv: astro-ph/9809321 . Bibcode:1999ApJ...513..805L. doi:10.1086/306900. S2CID   8914218.
  16. Li, H.; Finn, J. M.; Lovelace, R. V. E.; Colgate, S. A. (20 April 2000). "Rossby Wave Instability of Thin Accretion Disks. II. Detailed Linear Theory". The Astrophysical Journal. 533 (2): 1023–1034. arXiv: astro-ph/9907279 . Bibcode:2000ApJ...533.1023L. doi:10.1086/308693. S2CID   119382697.
  17. Petoukhov, Vladimir; Rahmstorf, Stefan; Petri, Stefan; Schellnhuber, Hans Joachim (2 April 2013). "Quasiresonant amplification of planetary waves and recent Northern Hemisphere weather extremes". Proceedings of the National Academy of Sciences of the United States of America. 110 (14): 5336–5341. Bibcode:2013PNAS..110.5336P. doi: 10.1073/pnas.1222000110 . PMC   3619331 . PMID   23457264.
  18. Mann, Michael E.; Rahmstorf, Stefan; Kornhuber, Kai; Steinman, Byron A.; Miller, Sonya K.; Coumou, Dim (30 May 2017). "Influence of Anthropogenic Climate Change on Planetary Wave Resonance and Extreme Weather Events". Scientific Reports. 7 (1): 45242. Bibcode:2017NatSR...745242M. doi:10.1038/srep45242. PMC   5366916 . PMID   28345645.

Bibliography