Firehose instability

Last updated
Fig. 1. The firehose instability in an N-body simulation of a prolate elliptical galaxy. Time progresses top-down, from upper left to lower right. Initially, the long-to-short axis ratio of the galaxy is 10:1. After the instability has run its course, the axis ratio is approximately 3:1. Note the boxy shape of the final galaxy, similar to the shapes of bars observed in many spiral galaxies. FirehoseNbody.png
Fig. 1. The firehose instability in an N-body simulation of a prolate elliptical galaxy. Time progresses top–down, from upper left to lower right. Initially, the long-to-short axis ratio of the galaxy is 10:1. After the instability has run its course, the axis ratio is approximately 3:1. Note the boxy shape of the final galaxy, similar to the shapes of bars observed in many spiral galaxies.

The firehose instability (or hose-pipe instability) is a dynamical instability of thin or elongated galaxies. The instability causes the galaxy to buckle or bend in a direction perpendicular to its long axis. After the instability has run its course, the galaxy is less elongated (i.e. rounder) than before. Any sufficiently thin stellar system, in which some component of the internal velocity is in the form of random or counter-streaming motions (as opposed to rotation), is subject to the instability.

Contents

The firehose instability is probably responsible for the fact that elliptical galaxies and dark matter haloes never have axis ratios more extreme than about 3:1, since this is roughly the axis ratio at which the instability sets in. [1] It may also play a role in the formation of barred spiral galaxies, by causing the bar to thicken in the direction perpendicular to the galaxy disk. [2]

The firehose instability derives its name from a similar instability in magnetized plasmas. [3] However, from a dynamical point of view, a better analogy is with the Kelvin–Helmholtz instability, [4] or with beads sliding along an oscillating string. [5]

Stability analysis: sheets and wires

The firehose instability can be analyzed exactly in the case of an infinitely thin, self-gravitating sheet of stars. [4] If the sheet experiences a small displacement in the direction, the vertical acceleration for stars of velocity as they move around the bend is

provided the bend is small enough that the horizontal velocity is unaffected. Averaged over all stars at , this acceleration must equal the gravitational restoring force per unit mass . In a frame chosen such that the mean streaming motions are zero, this relation becomes

where is the horizontal velocity dispersion in that frame.

For a perturbation of the form

the gravitational restoring force is

where is the surface mass density. The dispersion relation for a thin self-gravitating sheet is then [4]

The first term, which arises from the perturbed gravity, is stabilizing, while the second term, due to the centrifugal force that the stars exert on the sheet, is destabilizing.

For sufficiently long wavelengths:

the gravitational restoring force dominates, and the sheet is stable; while at short wavelengths the sheet is unstable. The firehose instability is precisely complementary, in this sense, to the Jeans instability in the plane, which is stabilized at short wavelengths, . [6]

Fig. 2. Unstable eigenmodes of a one-dimensional (prolate) galaxy. Growth rates are given at the left. FirehoseEvenModes.png
Fig. 2. Unstable eigenmodes of a one-dimensional (prolate) galaxy. Growth rates are given at the left.

A similar analysis can be carried out for a galaxy that is idealized as a one-dimensional wire, with density that varies along the axis. [7] This is a simple model of a (prolate) elliptical galaxy. Some unstable eigenmodes are shown in Figure 2 at the left.

Stability analysis: finite-thickness galaxies

At wavelengths shorter than the actual vertical thickness of a galaxy, the bending is stabilized. The reason is that stars in a finite-thickness galaxy oscillate vertically with an unperturbed frequency ; like any oscillator, the phase of the star's response to the imposed bending depends entirely on whether the forcing frequency is greater than or less than its natural frequency. If for most stars, the overall density response to the perturbation will produce a gravitational potential opposite to that imposed by the bend and the disturbance will be damped. [8] These arguments imply that a sufficiently thick galaxy (with low ) will be stable to bending at all wavelengths, both short and long.

Analysis of the linear normal modes of a finite-thickness slab shows that bending is indeed stabilized when the ratio of vertical to horizontal velocity dispersions exceeds about 0.3. [4] [9] Since the elongation of a stellar system with this anisotropy is approximately 15:1 — much more extreme than observed in real galaxies — bending instabilities were believed for many years to be of little importance. However, Fridman & Polyachenko showed [1] that the critical axis ratio for stability of homogeneous (constant-density) oblate and prolate spheroids was roughly 3:1, not 15:1 as implied by the infinite slab, and Merritt & Hernquist [7] found a similar result in an N-body study of inhomogeneous prolate spheroids (Fig. 1).

The discrepancy was resolved in 1994. [8] The gravitational restoring force from a bend is substantially weaker in finite or inhomogeneous galaxies than in infinite sheets and slabs, since there is less matter at large distances to contribute to the restoring force. As a result, the long-wavelength modes are not stabilized by gravity, as implied by the dispersion relation derived above. In these more realistic models, a typical star feels a vertical forcing frequency from a long-wavelength bend that is roughly twice the frequency of its unperturbed orbital motion along the long axis. Stability to global bending modes then requires that this forcing frequency be greater than , the frequency of orbital motion parallel to the short axis. The resulting (approximate) condition

predicts stability for homogeneous prolate spheroids rounder than 2.94:1, in excellent agreement with the normal-mode calculations of Fridman & Polyachenko [1] and with N-body simulations of homogeneous oblate [10] and inhomogeneous prolate [7] galaxies.

The situation for disk galaxies is more complicated, since the shapes of the dominant modes depend on whether the internal velocities are azimuthally or radially biased. In oblate galaxies with radially-elongated velocity ellipsoids, arguments similar to those given above suggest that an axis ratio of roughly 3:1 is again close to critical, in agreement with N-body simulations for thickened disks. [11] If the stellar velocities are azimuthally biased, the orbits are approximately circular and so the dominant modes are angular (corrugation) modes, . The approximate condition for stability becomes

with the circular orbital frequency.

Importance

The firehose instability is believed to play an important role in determining the structure of both spiral and elliptical galaxies and of dark matter haloes.

See also

Related Research Articles

<span class="mw-page-title-main">Nonlinear optics</span> Branch of physics

Nonlinear optics (NLO) is the branch of optics that describes the behaviour of light in nonlinear media, that is, media in which the polarization density P responds non-linearly to the electric field E of the light. The non-linearity is typically observed only at very high light intensities (when the electric field of the light is >108 V/m and thus comparable to the atomic electric field of ~1011 V/m) such as those provided by lasers. Above the Schwinger limit, the vacuum itself is expected to become nonlinear. In nonlinear optics, the superposition principle no longer holds.

<span class="mw-page-title-main">Wave</span> Repeated oscillation around equilibrium

In physics, mathematics, engineering, and related fields, a wave is a propagating dynamic disturbance of one or more quantities. Waves can be periodic, in which case those quantities oscillate repeatedly about an equilibrium (resting) value at some frequency. When the entire waveform moves in one direction, it is said to be a traveling wave; by contrast, a pair of superimposed periodic waves traveling in opposite directions makes a standing wave. In a standing wave, the amplitude of vibration has nulls at some positions where the wave amplitude appears smaller or even zero. Waves are often described by a wave equation or a one-way wave equation for single wave propagation in a defined direction.

<span class="mw-page-title-main">Cutoff frequency</span> Frequency response boundary

In physics and electrical engineering, a cutoff frequency, corner frequency, or break frequency is a boundary in a system's frequency response at which energy flowing through the system begins to be reduced rather than passing through.

<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.

In fluid dynamics, Stokes' law is an empirical law for the frictional force – also called drag force – exerted on spherical objects with very small Reynolds numbers in a viscous fluid. It was derived by George Gabriel Stokes in 1851 by solving the Stokes flow limit for small Reynolds numbers of the Navier–Stokes equations.

<span class="mw-page-title-main">Stellar dynamics</span>

Stellar dynamics is the branch of astrophysics which describes in a statistical way the collective motions of stars subject to their mutual gravity. The essential difference from celestial mechanics is that the number of body

Plasma oscillations, also known as Langmuir waves, are rapid oscillations of the electron density in conducting media such as plasmas or metals in the ultraviolet region. The oscillations can be described as an instability in the dielectric function of a free electron gas. The frequency depends only weakly on the wavelength of the oscillation. The quasiparticle resulting from the quantization of these oscillations is the plasmon.

<span class="mw-page-title-main">Bending</span> Strain caused by an external load

In applied mechanics, bending characterizes the behavior of a slender structural element subjected to an external load applied perpendicularly to a longitudinal axis of the element.

<span class="mw-page-title-main">Two-state quantum system</span> Simple quantum mechanical system

In quantum mechanics, a two-state system is a quantum system that can exist in any quantum superposition of two independent quantum states. The Hilbert space describing such a system is two-dimensional. Therefore, a complete basis spanning the space will consist of two independent states. Any two-state system can also be seen as a qubit.

<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.

<span class="mw-page-title-main">Perfectly matched layer</span>

A perfectly matched layer (PML) is an artificial absorbing layer for wave equations, commonly used to truncate computational regions in numerical methods to simulate problems with open boundaries, especially in the FDTD and FE methods. The key property of a PML that distinguishes it from an ordinary absorbing material is that it is designed so that waves incident upon the PML from a non-PML medium do not reflect at the interface—this property allows the PML to strongly absorb outgoing waves from the interior of a computational region without reflecting them back into the interior.

<span class="mw-page-title-main">Inertial wave</span>

Inertial waves, also known as inertial oscillations, are a type of mechanical wave possible in rotating fluids. Unlike surface gravity waves commonly seen at the beach or in the bathtub, inertial waves flow through the interior of the fluid, not at the surface. Like any other kind of wave, an inertial wave is caused by a restoring force and characterized by its wavelength and frequency. Because the restoring force for inertial waves is the Coriolis force, their wavelengths and frequencies are related in a peculiar way. Inertial waves are transverse. Most commonly they are observed in atmospheres, oceans, lakes, and laboratory experiments. Rossby waves, geostrophic currents, and geostrophic winds are examples of inertial waves. Inertial waves are also likely to exist in the molten core of the rotating Earth.

Atmospheric tides are global-scale periodic oscillations of the atmosphere. In many ways they are analogous to ocean tides. Atmospheric tides can be excited by:

In general relativity, Lense–Thirring precession or the Lense–Thirring effect is a relativistic correction to the precession of a gyroscope near a large rotating mass such as the Earth. It is a gravitomagnetic frame-dragging effect. It is a prediction of general relativity consisting of secular precessions of the longitude of the ascending node and the argument of pericenter of a test particle freely orbiting a central spinning mass endowed with angular momentum .

The McCumber relation is a relationship between the effective cross-sections of absorption and emission of light in the physics of solid-state lasers. It is named after Dean McCumber, who proposed the relationship in 1964.

<span class="mw-page-title-main">Stokes wave</span> Nonlinear and periodic surface wave on an inviscid fluid layer of constant mean depth

In fluid dynamics, a Stokes wave is a nonlinear and periodic surface wave on an inviscid fluid layer of constant mean depth. This type of modelling has its origins in the mid 19th century when Sir George Stokes – using a perturbation series approach, now known as the Stokes expansion – obtained approximate solutions for nonlinear wave motion.

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.

In the fields of nonlinear optics and fluid dynamics, modulational instability or sideband instability is a phenomenon whereby deviations from a periodic waveform are reinforced by nonlinearity, leading to the generation of spectral-sidebands and the eventual breakup of the waveform into a train of pulses.

The Darrieus–Landau instability or hydrodynamic instability is an instrinsic flame instability that occurs in premixed flames, caused by the density variation due to the thermal expansion of the gas produced by the combustion process. In simple terms, the stability inquires whether a steadily propagating plane sheet with a discontinuous jump in density is stable or not. It was predicted independently by Georges Jean Marie Darrieus and Lev Landau.

In the physics of continuous media, spatial dispersion is a phenomenon where material parameters such as permittivity or conductivity have dependence on wavevector. Normally, such a dependence is assumed to be absent for simplicity, however spatial dispersion exists to varying degrees in all materials.

References

  1. 1 2 3 Fridman, A. M.; Polyachenko, V. L. (1984), Physics of Gravitating Systems. II — Nonlinear collective processes: Nonlinear waves, solitons, collisionless shocks, turbulence. Astrophysical applications, Berlin: Springer, ISBN   978-0-387-13103-0
  2. 1 2 Raha, N.; Sellwood, J. A.; James, R. A.; Kahn, F. A. (1991), "A dynamical instability of bars in disk galaxies", Nature, 352 (6334): 411–412, Bibcode:1991Natur.352..411R, doi:10.1038/352411a0, S2CID   4258274
  3. Parker, E. N. (1958), "Dynamical Instability in an Anisotropic Ionized Gas of Low Density", Physical Review, 109 (6): 1874–1876, Bibcode:1958PhRv..109.1874P, doi:10.1103/PhysRev.109.1874
  4. 1 2 3 4 Toomre, A. (1966), "A Kelvin–Helmholtz Instability", Notes from the Geophysical Fluid Dynamics Summer Study Program, Woods Hole Oceanographic Inst.: 111–114
  5. In spite of its name, the firehose instability is not related dynamically to the oscillatory motion of a hose spewing water from its nozzle.
  6. Kulsrud, R. M.; Mark, J. W. K.; Caruso, A. (1971), "The Hose-Pipe Instability in Stellar Systems", Astrophysics and Space Science, 14 (1): 52–55, Bibcode:1971Ap&SS..14...52K, doi:10.1007/BF00649194, S2CID   120864161.
  7. 1 2 3 Merritt, D.; Hernquist, L. (1991), "Stability of Nonrotating Stellar Systems", The Astrophysical Journal, 376: 439–457, Bibcode:1991ApJ...376..439M, doi: 10.1086/170293 .
  8. 1 2 Merritt, D.; Sellwood, J. (1994), "Bending Instabilities of Stellar Systems", The Astrophysical Journal, 425: 551–567, Bibcode:1994ApJ...425..551M, doi:10.1086/174005
  9. Araki, S. (1985). "A Theoretical Study of the Stability of Disk Galaxies and Planetary Rings. PhD Thesis, MIT". OCLC   13915550.{{cite journal}}: Cite journal requires |journal= (help)
  10. Jessop, C. M.; Duncan, M. J.; Levison, H. F. (1997), "Bending Instabilities in Homogenous Oblate Spheroidal Galaxy Models", The Astrophysical Journal, 489 (1): 49–62, Bibcode:1997ApJ...489...49J, doi: 10.1086/304751 , S2CID   120230527
  11. Sellwood, J.; Merritt, D. (1994), "Instabilities of counterrotating stellar disks", The Astrophysical Journal, 425: 530–550, Bibcode:1994ApJ...425..530S, doi:10.1086/174004
  12. Bett, P.; et al. (2007), "The spin and shape of dark matter haloes in the Millennium simulation of a Λ cold dark matter universe", Monthly Notices of the Royal Astronomical Society , 376 (1): 215–232, arXiv: astro-ph/0608607 , Bibcode:2007MNRAS.376..215B, doi:10.1111/j.1365-2966.2007.11432.x, S2CID   119466166
  13. 1 2 Combes, F.; et al. (1990), "Box and peanut shapes generated by stellar bars", Astronomy and Astrophysics, 233: 82–95, Bibcode:1990A&A...233...82C
  14. Revaz, Y.; Pfenniger, D. (2004), "Bending instabilities at the origin of persistent warps: A new constraint on dark matter halos", Astronomy and Astrophysics, 425: 67–76, arXiv: astro-ph/0406339 , Bibcode:2004A&A...425...67R, doi:10.1051/0004-6361:20041386, S2CID   5424745