Issue |
A&A
Volume 521, October 2010
|
|
---|---|---|
Article Number | A46 | |
Number of page(s) | 16 | |
Section | Planets and planetary systems | |
DOI | https://doi.org/10.1051/0004-6361/201014412 | |
Published online | 19 October 2010 |
Fine-scale density wave structure of Saturn's rings: A hydrodynamic theory
E. Griv - M. Gedalin
Department of Physics, Ben-Gurion University of the Negev, PO Box 653, Beer-Sheva 84105, Israel
Received 12 March 2010 / Accepted 13 July 2010
Abstract
Aims. We examine the linear stability of the Saturnian ring
disk of mutually gravitating and physically colliding particles with
special emphasis on its fine-scale m
density wave structure, that is, almost regularly spaced, aligned
cylindric density enhancements and optically-thin zones with the width
and the spacing between them of roughly several tens particle
diameters.
Methods. We analyze the Jeans' instabilities of gravity
perturbations (e.g., those produced by a spontaneous disturbance)
analytically by using the Navier-Stokes dynamical equations of a
compressible fluid. The theory is not restricted by any assumptions
about the thickness of the system. We consider a simple model of the
system consisting of a three-dimensional ring disk that is weakly
inhomogeneous and whose structure is analyzed by making a horizontally
local short-wave approximation.
Results. We demonstrate that the disk is probably unstable and
that gravity perturbations grow effectively within a few orbital
periods. We find that self-gravitation plays a key role in the
formation of the fine structure. The predictions of the theory are
compared with observations of Saturn's rings by the Cassini
spacecraft and are found to be in good agreement. In particular,
it appears very likely that some of the quasi-periodic
microstructures observed in Saturn's A and B rings -
both axisymmetric and nonaxisymmetric ones - are manifestations of
these effects. We argue that the quasi-periodic density enhancements
revealed in Cassini data are flattened structures, with a
height to width ratio of about 0.3. One should analyze
high-resolution of the order of 10 m data acquired for the A
and B rings (and probably C ring as well) to confirm
this prediction. We also show that the gravitational instability is a
potential cluster-forming mechanism leading to the formation of porous
100-m-diameter moonlets of preferred mass
g each embedded in the outer A ring, although this has yet to be directly measured.
Key words: planets and satellites: general - planets and satellites: rings - planets and satellites: individual: Saturn - instabilities - hydrodynamics
1 Introduction
The dynamics of highly flattened, rapidly rotating gravitating systems have been studied quite thoroughly. This research has attempted to explain the origin of spiral and ring structures in galaxies, in disks around supermassive black holes, in protostellar and protoplanetary clouds, in Saturn's ring system, in the narrow and widely separated ring system of Uranus, etc. A primary theme has been to analyze the perturbation dynamics of these systems, in both linear and nonlinear regimes. We refer to Lin & Lau (1979), Shu (1984), Papaloizou & Lin (1995), Lin & Papaloizou (1996), Bertin & Lin (1996), Bertin (2000), and Yuan (2002) for reviews of the theory and its applications. It has been shown that nonuniformly rotating self-gravitating astrophysical disk configurations are highly dynamic and subject to various collective instabilities of gravity perturbations (e.g., those produced by a spontaneous disturbance or, in rare cases, a companion system). This is because their evolution is primarily driven by angular momentum redistribution. The system may then fall toward the lower potential (orbital) energy configuration and use the energy so gained to increase its coarse-grained entropy (Lynden-Bell & Kalnajs 1972; Griv & Gedalin 2004; Griv et al. 2008).
The theoretical studies of Maxwell (1859)
demonstrated that the rings around Saturn might not be either solid or
liquid, but rather a swarm of millions of individual particles rotating
in separate concentric orbits at different speeds. A modern very
popular model of the particles in Saturn's rings is a smooth ice
sphere, whose restitution coefficient is quite high and decreases as
the collision velocity increases (e.g., Kerr 1985). Ring particles are in circular orbital motion in Saturn's equatorial plane, at circular velocities
km s-1 and relative (random) velocities
cm s-1.
The Saturnian ring disk consists predominantly of water-ice particles
ranging between about 1 cm and 5 m in radius a
with a distance between them of about a few meters; the bulk
of Saturn's ring mass is assumed to reside in individual particles in
the 1-3 m diameter range (Shu 1984; Zebker et al. 1985; French & Nicholson 2000; Cuzzi et al. 2002, 2009; Esposito 2002; Griv & Gedalin 2003).
Numerical simulations have indicated that the larger particles are
nearly confined to a monolayer, with the smaller particles filling the
spaces between the larger particles (Zebker et al. 1985; Altobelli et al. 2008). The comparison of ground-based visible data with C ASSINI CIRS observations also indicated that the A ring is a monolayer (Leyrat et al. 2008a; Ferrari et al. 2009).
V OYAGER flybys of Saturn have revealed that the disk about the planet is not simply divided into several main bands, the classical A-C rings, but that the entire disk assembly is indeed subdivided into a huge number of fine thread-like rings with the appearance of record-grooves (Smith et al. 1981, 1982; Lane et al. 1982; Stone & Miner 1982; Brahic 2001, Figs. 7-9 therein; Cuzzi et al. 2002, Fig. 2b therein; Esposito 2002, Fig. 5 therein). It is important that the V OYAGER's photopolarimeter PPS data detected some indirect evidence of ``finest'' structuring in the densest central parts of the opaque Saturn's B ring down to the 100 m length scale (Esposito et al. 1983; Showalter & Nicholson 1990; Nicholson & Dones 1991; Horne & Cuzzi 1996; Esposito 2002, p. 1751). However, below a length scale of a few kilometers, the V OYAGER's PPS data is too noisy to extract information about the structure: the finest structure observed by PPS is accurately descibed by models of statistical noise combined with stochastic variations resulting from large particles or clumps of particles (Showalter & Nicholson 1990). It was suggested numerically by Salo (1992, 1995) and analytically by Griv (1998, 2005a), Griv et al. (2000, 2003a,b), and Griv & Gedalin (2003, 2005) that far more precise C ASSINI spacecraft observations would help us to resolve the question.
The fine-scale of the order of 100 m or even less quasi-periodic
density structure of the Saturnian brightest A and B rings
has indeed been discovered by C ASSINI science images (Porco et al. 2005, Figs. 5A and F therein), and then C ASSINI UVIS and radio occultations (Colwell et al. 2006, 2007; Thomson et al. 2007). C ASSINI VIMS observations (Hedman et al. 2007) and infrared observations of Saturn's rings by C ASSINI CIRS (Leyrat et al. 2008a; Ferrari et al. 2009)
also detected fine structure in the A ring. Opaque particle clumps
called self-gravity wakes are separated by nearly empty, optically-thin
gaps. Interestingly, both nonaxisymmetric structures that have a
characteristic trailing orientation of
relative to local direction
of orbital motion (Colwell et al. 2006, 2007; Hedman et al. 2007; Leyrat et al. 2008a; Ferrari et al. 2009) and azimuthally symmetric structures with the spatial orientation
(Colwell et al. 2007, p. 141; Thomson et al. 2007, p. L24203) were found. These structures are found to have an average height-to-spacing ratio
and a width to spacing ratio of
.
Gaps between wakes, which are filled with particles, have a low optical depth
.
To reiterate, C ASSINI
results show that in both the A and B rings, axisymmetric
microstructure, characterized by a periodic radial variation in optical
depth, coexists with nonaxisymmetric microstructure. The
spacing between these structures vary from 30 m
to 250 m, depending on the location in the rings. Nicholson
et al. (2008) noted that the
quasi-periodic microstructure in the inner B ring appears to be
unchanged in amplitude and phase over an interval of 24 years.
Both spacecraft missions have shown that these relatively large
``irregular variations'' in optical depth are not associated with any
resonances with known satellites. On a small scale, the fine variations
have surprizingly been observed to undergo variation and oscillations
with time and ring longitude (Lane et al. 1982; Smith et al. 1982).
The latter indicates that the irregular variations are probably wave
phenomena, and different instabilities of small-amplitude gravity
perturbations may play important roles in ring dynamics. A wavelet
analysis of the structure of A-C rings has already shown that
the rings exhibit various wave perturbations, which weakly interact
with each other (Postnikov & Loskutov 2007).
One concludes that at the sub-km level the dense A and
B rings are dominated by an elongated, quasi-periodic
microstructures also known as self-gravity wakes in the literature (see
Nicholson & Hedman 2010, for a discussion)
.
In our model, Saturn's rings consist of primarily larger than 1 cm
size identical, almost elastically colliding, and gravitating
particles. The model formation is thought to start with partially
inelastically colliding particle settling onto the central plane of a
rotating cloud to form a thin and relatively dense disk around the
plane. Because of inelastic physical impacts, the disk radiates
heat from its surface and, therefore, cools becoming thinner and
thinner. The mere existence of flattened rings implies dissipation by
means of partially inelastic impacts, which takes some 106 yr
to complete. Subsequently, the local gravitational instability causes
the disk on attaining a certain critical thickness, which is small in
comparison with the outer radius of the system R (and, correspondingly, very low temperature), to disintegrate spontaneously into a number of separate rings and spirals.
Thus, we consider the gravitationally unstable rings to be an annular
disk with concentric circles and spiral local maxima and minima in both
density and brightness. It is natural to assume that the growth
rate of the gravity perturbations of an originally weakly unstable
cloud was rather low. Therefore, we investigate perturbations that are
relatively small
compared to equilibrium (though of finite amplitude).
2 Formulation of the problem
As for high-temperature plasmas, a system of particles in Saturn's rings exhibits collectively unstable modes of motions. Because of its long-range Newtonian forces, a self-gravitating medium (a particulate ``gas'', say) possesses collective motions in which all the particles of the system participate. These properties should be manifested in the behavior of small gravity perturbations arising against the equilibrium background. ``Collective processes are completely analogous to two-body collisions, except that one particle collides with many which are collected together by some coherent process such as a wave'' (Kulsrud 1972, p. 338). Thus, relaxation in particulate systems can occur without ordinary collisions by means of the influence of collective motions of the particulate gas on the particle distribution (Kulsrud 1972; Sagdeev & Galeev 1969; Galeev & Sagdeev 1983; Griv et al. 2006b). A cloud of ``single charged'' particles of Saturn's rings can be unstable against azimuthal and axial perturbations, i.e., in general, inconsistent with a being in thermodynamic equilibrium.
The number of mechanisms producing the ubiquitous fine-scale structure of Saturn's rings has grown rapidly in the past 25 years. In a review by Griv & Gedalin (2003), seven mechanisms are listed. However, along with this growth in the number of known mechanisms, there has been a growth in our understanding that a universal mechanism may generate the density structure in all regions of Saturn's main rings. In the present paper of the series, we present the subject of Saturn's rings instabilities in such a way as to emphasize that this universal mechanism exists. We regard this microstructure of rings about Saturn as a wave pattern, which does not remain stationary in a frame of reference rotating around the planet at a proper speed, excited as a result of Jeans' nonresonant (or algebraic) gravitational instability. The nonaxisymmetric wave structure structure rotates uniformly, although the material rotates differentially and the spirals (and rings) consist of different material at different times.
On the other hand, there are regular ringlet complexes in the A-C rings connected to resonances with external satellites, including Lindblad horizontal and vertical resonances (Holberg et al. 1982; Lissauer & Cuzzi 1982).
The structures associated with this kind of resonances are directly
observed as the so-called spiral or bending wave trains (Shu
et al. 1983; Shu 1984; Rosen et al. 1991; Yuan 2002, Fig. 3 therein; Esposito 2002, p. 1752; Porco et al. 2005, Figs. 5G and H therein; Sicardy 2005, p. 463; Tiscareno et al. 2006a, 2007; Colwell et al. 2009).
The trains are the resonantly excited density waves that decay as they
propagate away from the resonances. Second-order waves were also
clearly resolved in the C ASSINI data (Tiscareno et al. 2006a).
The waves can be used as diagnostics to obtain fundamental physical
parameters that characterize the dynamical state of the ring such as
mass, thickness, and collision velocities (Shu et al. 1983, 1985; Yuan 2002). Hedman et al. (2009)
presented images of resonant structures in Saturn's faint G ring,
the inner Roche Division, and the middle D ring. Most of the
structures in Saturn's A and B rings, however, do not
correspond to resonances with known satellites: wave trains associated
with known resonances cover less than
of the radial extent of the A and B rings (Goldreich & Tremaine 1982;
Horn & Cuzzi 1996). The study of these regular wave ringlets is
beyond the scope of the present paper. We also do not consider few
truly isolated ringlets with adjacent empty gaps, located in the
low-density C ring, in the inner B ring, in the
A ring, and in the Cassini Division,
resembling those of Uranus (Porco 1990; Esposito 2002,
p. 1756). Many of the narrow ringlets with typical widths of
a few tens of kilometers and extremely sharp edges are found in
the isolated resonance locations of different satellites (Rosen
et al. 1991).
An external satellite takes angular momentum from the particles of
a disk, and the resultant angular momentum transfer can open gaps or
terminate rings (e.g., Griv 2007a).
It is convenient to divide instabilities into two broad classes: (a) macroscopic and (b) microscopic. The macroscopic, or hydrodynamic, instabilities imply the displacement of macroscopic portions of rings - all the particles in a given macroscopic volume execute the same average motion. The gasdynamic equations and the continuity equation can be used for the theoretical analysis of this class of instabilities. The microscopic, or kinetic, instabilities can be defined as those for which the differences in the motion of different particles in the same volume are important. The Boltzmann equations are necessary for the analysis of these instabilities. Our aim is to present the linear theory of macroscopic Jeans instabilities of Saturn's rings. The present instability theory assumes that the departure of the system from dynamical equilibrium is infinitesimal and then ask whether this infinitesimal departure grows or decays. Notice that although the linear theory does not establish the amplitude of the perturbations, it does yield values of their dispersion properties and by means of stability criteria it can determine some of the equilibrium parameters of the system under study.
The Jeans instability sets in when the destabilizing effect of
the self-gravity in the disk exceeds the combined restoring action of
the pressure and Coriolis forces. The wave propagation is a process of
rotation as a solid about the center at a fixed phase velocity, despite
the general differential rotation of the system; the alternating
density enhancements and relatively rarefied zones consist of different
material at different times.
The classical Jeans instability of gravity disturbances is one of the
most frequent and important instabilities in the stellar and planetary
cosmogony, and galactic dynamics. The term gravitational instability,
as introduced by Jeans (1928),
deals with the question of whether initial density fluctuations will
either be amplified or die down. Jeans instability identifies
nonresonant
instabilities of gravity fluctuations associated with almost
aperiodically growing accumulations of mass, and the dynamics of Jeans
perturbations can be characterized as a fluidlike wave-particle
interaction. In other words, the instability associated with
departures of macroscopic quantities from the dynamical equilibrium is
hydrodynamical in nature and has nothing to do with any explicit
resonant
effects, where
is the oscillation frequency, k is the wavenumber, and v is the particle's velocity. Following Lin & Shu (e.g., Lin et al. 1969),
a relatively simple hydrodynamical model can be used to
investigate the instability (Lau & Bertin 1978; Lin & Lau 1979; Morozov 1985; Montenegro et al. 1999; Griv
2006). Thus a kinetic description (e.g., Bertin 1980; Griv et al. 2002) yields results that are almost no different from those obtained hydrodynamically. Ginzburg et al. (1972)
first examined the possible gravitational instability of Saturn's rings
of the type discussed by Lin & Shu in the context of the formation
of spiral arms of normal galaxies.
The difficulties in a satisfactory understanding of the dynamics of particulate systems are due to the well known fact that systems of N gravitationally interacting particles are naturally inhomogeneous, because gravitational forces between gravitational ``charges'' (which are always ``charges of the same sign'') are always attractive. The density of the gravitating disk decreases towards its periphery. This inherent inhomogeneity character is the origin of some mathematical problems. For instance, in the case of spiral galaxies one has to deal with complex boundary problems while in a plasma the model of a homogeneous infinitely extended background already allows one to obtain many familiar effects. Following plasma physics, however, the problem may be simplified by considering the so-called Wenzel-Kramer-Brillouin (WKB), or short-wave local approximation. Under the local consideration (in the vicinity of a given point) of perturbations with scales small compared to a characteristic linear dimension R of the system, one may assume parameters of the stationary state equal to its values in a given point. The local approach applied to the gravitational instability and to many types of instabilities in an inhomogeneous self-gravitating systems is valuable, in particular for understanding the physics of the phenomenon considered, since it leads to rather simple analytical results.
In closing a section, one usually chooses a model of Saturn's
rings consisting of an infinitesimally thin disk of particles with the
surface density equal to the projection of the full mass density on the
plane perpendicular to the rotating axis, i.e., a disk,
the equilibrium thickness 2h of which is many times smaller than the perturbation radial wavelength
,
where
is
the radial wave number. It has been stated that this accuracy is
sufficient for the discussion of oscillation modes in Saturn's rings of
small but finite disk thickness. As is known, the standard
Lin-Shu dispersion relation for planar density waves hypothesized to
propagate in galaxies (Lin & Shu 1966; Lin et al. 1969; Shu 1970) has been slightly improved in either a heuristic manner (Lin & Shu 1968; see also Toomre 1964; and Safronov 1980) or by introducing an approximate reduction factor providing the correction for finite thickness (Vandervoort 1970). We refer to Shu (1984) for a review of the problem. The method employed by Vandervoort (1970) to solve the collisionless Boltzmann equation when normal to the equatorial plane, z-motions are taken into additional account is based on the existence of an adiabatic invariant Jz, whose approximate constancy characterizes the vertical motion of disk particles. Romeo (1992) already stated that this is a characteristic of only highly flattened, rotationally supported astrophysical disks, where the frequency of the free epicyclic oscillations in the z direction
is large
compared to the frequency of the epicyclic motion in the symmetry
plane, which in turn is generally of the same order as the pattern
frequency of spiral waves. We present a more rigorous investigation of
the effect of finite thickness (or the particle velocity spread in
the z direction)
on the propagation of collective oscillations. A different
approach from the rest is taken: (1) to study the effect of
thickness on even-parity gravity perturbations of the kind studied by
Lin & Shu (Fig. 1b), which do not distort the plane of a disk, by considering the arbitrary but not only highly flattened, rapidly rotating (Vandervoort 1970, p. 94; Shu 1984,
p. 557; Romeo 1992, p. 311) system and (2) to treat the odd-parity perturbations of the bending type (Fig. 1c), which do distort the disk's plane. The problem is formulated in the same way as in plasma theory (Arsenin 1967; Harris 1968; Alexandrov et al. 1984, p. 313). This approach is introduced here for the first time in an astrophysical context.
![]() |
Figure 1: Sketch of perturbations of a three-dimensional disk. In a) a section of the disk is shown edge-on. In b) a mode of even symmetry with respect to the equatorial plane, or an even-parity Jeans-type perturbation, is shown (the dashed line). In c) a mode of odd symmetry with respect to the equatorial plane, or an odd bending-type perturbation, is illustrated (the dashed line). |
Open with DEXTER |
3 Basic equations
An extended flat Saturnian ring disk of identical, mutually
gravitating, and elastically colliding particles orbiting the planet is
studied, with
being Saturn's mass,
being the mass of the disk, and
.
That h/R
is small means that the disk considered by us is rather dynamically
cold and that the pressure gradient in it is much smaller than both the
gravitational and the centrifugal forces. The model assumes ring
particles to be confined by a strong Saturnian gravitational field.
To formulate the problem, we must include effects of
self-gravitation, pressure gradients, and viscosity. The evolution of
the system can be described by the basic equations of viscous fluid
dynamics, or Navier-Stokes model motion equations
and the continuity equation
where














and

Thus, at the outset we adopt an isothermal fluid model for Saturn's rings, identifying the velocity dispersion of identical particles in the rings with the sound speed in our model. In a self-consistent problem, Eqs. (1)-(3) must be solved simultaneously with the Poisson equation
Owing to their velocity dispersion, ring particles experience collisions. In its simplest form, the nature of the disk viscosity is physical collisions and the frequency of collisions is obviously



(Jeffreys 1947; Cook & Franklin 1964; Goldreich & Tremaine 1982; Bridges et al. 1984; Stewart et al. 1984; Araki 1991). In the three-dimensional disk, the characteristic frequency of collisions is indeed
where




(bearing in mind that










In Appendix A below, we show that the system of Eqs. (1)-(4) admits a solution in the form of self-gravity
density waves, which were first studied by Lin & Shu (Lin & Shu 1966, 1968; Lin et al. 1969; Shu 1970; Lin & Lau 1979) on larger scales as they exist in galactic disks. These fully self-consistent Lin-Shu density waves (normal modes) are not to be confused, of course, with Julian & Toomre's (1966) perturbed or ``forced'' density wave proposal explored by Colombo et al. (1976), Franklin & Colombo (1978), Lumme & Irvine (1979), and Karttunen (1983). Julian & Toomre (1966)
worked out the response of a disk - a calculation of the
polarization cloud - when forced by an orbiting mass or clump,
such as a giant molecular cloud in spiral galaxies.
Accordingly, the integrated mass of the polarization cloud (``wake'')
may be several times the mass of the original clump (Julian &
Toomre 1966, Fig. 7 therein). According to Colombo et al. (1976),
in Saturn's rings ``large particles will force intense trailing
density wakes'' and this Julian-Toomre-type mechamism for producing the
azimuthal brightness variation ``requires the rings to contain
particles whose radii are considerably larger than average''.
In the spirit of Julian & Toomre (1966), to calculate the response of a disk to a point mass orbiting within it, one needs to replace the potential
in Eqs. (1)-(4) above by
,
where
is the imposed potential (Julian & Toomre 1966, p. 811). The results following from Eqs. (1)-(4) obviously critically depend on
;
the polarization cloud vanishes if
.
Unlike Julian & Toomre (1966), we develop the self-consistent theory of real instabilities of small spontaneous disturbances (the external potential
),
which develop effectively from noise to observable amplitudes in some
20 h in the main part of the system under study.
As for us, Julian & Toomre's (1966) ``forced'' instability of a system of almost identical particles cannot serve as the generator of quasi-regular spiral (and ring)
density waves with a characteristic pitch angle
in Saturn's
rings
.
The Lin-Shu-type density wave conception of microstructure adopted in
the present paper is less restrictive than the Julian-Toomre one. We
suggest that the preferred orientation of self-excited gravitationally
unstable density waves (Lin-Shu-type ``free'' density waves) produces
brightness azimuthal variations that have been observed
(see, e.g., Leyrat et al. 2008a,b, for an explanation).
3.1 Dynamical equilibrium
We now assume that
.
Let us consider a free particle orbiting in a
circle of radius r with angular speed
in the equatorial plane z=0 of the non-spherical planet with an associated gravitational potential
.
The planar equilibrium of a system is expressed by the condition
If this test particle is displaced by an arbitrary small amount, it will oscillate freely in the horizontal and vertical directions about the reference circular orbit with epicyclic frequency


In the epicyclic approximation, the motion of a particle is represented as in epicyclic motion along the small ellipse (epicycle) with a simultaneous circulation of the epicenter about the planetary center. Of course, the epicyclic approximation may be applied only when the true particle motion is nearly circular, as in planetary rings, and the collision frequency is small in comparison with







![]() |
Figure 2: Schematic model of the fine-scale density wave structure in Saturn's rings. Self-gravity density waves, which were first studied by Lin & Shu (Lin et al. 1969; Shu 1970), manifest themselves as evenly spaced elongated clusters of ring particles. Shown are both a three-dimensional distribution of particles a) and a distribution of particles in the (x,z)-plane b). |
Open with DEXTER |
The unperturbed disk is assumed to have no motion except for rotation.
The present theory suggests some perturbed radial, azimuthal, and
vertical motions of the fluid element distributed in the form of a
spiral-like flow field, which is a small correction to the basic
circular, equilibrium motion. In a plane perpendicular to the rotation
axis, the equilibrium motion is described by the equation
where






where cz is the sound speed along the normal to the plane direction. By considering the geometrically thin disk,


and

![[*]](/icons/foot_motif.png)
where




Since realistic three-dimensional Saturn's rings are extremely
difficult to treat, we consider an idealized model. First,
in Eq. (13) for a weakly spatially inhomogeneous disk we consider values of
and P0 at r=r0 (the local approximation); their distributions are assumed to be symmetric with respect to the z = 0 plane. Second, in a first step of our study and in the spirit of Mark (1971), Kulsrud et al. (1971), and Bertin & Casertano (1982), the density in the equilibrium state, along the z coordinate is regarded as uniform between two boundaries -h and +h, with a vacuum exterior (Fig. 1a).
In other words, following these authors, we consider here a slab
equilibrium with plane-parallel symmetry that is homogeneous in the
vertical, z-direction. Furthermore, we restrict ourselves to
consideration of only Lin-Shu unforced density waves, or ``heavy
sound,'' which are nothing but longitudinal compression waves in which
the self-gravitation of the fluctuations in density is taken into
account. These modes have wave vectors perpendicular to the axis of
rotation (Fig. 1b). N-body
experiments have already identified these longitudinal collective
motions for a system of mutually gravitating particles in computer
models of Saturn's rings, that is, the particle motion is
restricted to be almost parallel to the equatorial plane of the system
(Griv 2005a). The waves compress
existing material and cause both spiral and clump (``moonlets'')
formation. The modes of odd symmetry with respect to the equatorial
plane (Fig. 1c) deserve a
separate investigation. Thus, under the action of the gravitational
field of the planet, the particles move along almost closed
orbits. It is this equilibrium model of the Saturnian ring disk
that is examined for stability in the present investigation. The
schematic model of self-gravity Jeans-unstable density waves we
investigate is shown in Fig. 2.
4 Astronomical implications
In Appendix A, according to the local approximation of the WKB method it is shown that Saturn's ring disk rotating differentially could be unstable because of Jeans' gravitational instability. To suppress Jeans instabilities, a generalized local stability criterion must be satisfied, which indicates that in a differentially rotating disk, nonaxisymmetric perturbations are more unstable than axisymmetric ones. Equation (A.31) indicates the stabilizing influence of the finite thickness of a disk, which causes a shift in the threshold of instability toward a longer wavelength (and larger wavelength will include more mass). The etiology of this fluidlike instability is the nonresonant interaction of particles with Jeans-unstable gravity perturbations in a disk. It is similar to the instability of the bunching type in plasmas, e.g., pinch instabilities or a firehose instability. The nonresonant gravitational instability does not depend on the behavior of the particle distribution function in the neighborhood of a particular speed, but the determining factors of the instability are macroscopic parameters such as the velocity dispersion, mean density, and angular velocity of regular rotation. We now proceed to discuss the extent to which our results on the disk's stability can have a bearing on observable Saturn's rings.
4.1 Stabilizing effect of the finite thickness
The Saturnian ring disk is regarded as being at the threshold of instability. At the limit of gravitational stability
and we assume that the viscosity to have no essential effect on the
criterion of the gravitational instability. Substituting this
expression for h into Eq. (A.34), we estimate the stabilizing factor introduced by the finite thickness
.
Morozov (1980, 1981)
took into account the additional weak destabilizing effect of a density
inhomogeneity and a radial gradient in
velocity dispersion. The result is that the stabilizing effect
introduced by the finite thickness and the destabilizing effect
introduced by the inhomogeneity and the velocity gradient practically
cancel each other out, at least in the local stellar disk of
our own Galaxy. In practice, one can
neglect all these corrections. We expect therefore that in Saturn's
rings
(or Toomre's stability parameter
). Interestingly, about the same value of
bring the observations of Saturn's rings (Lane et al. 1982, p. 543).
4.2 Fine-scale structure
Thus, the modified Safronov-Toomre stability criterion given by Eq. (A.34)
is deduced, according to which the system is violently Jeans-unstable
against the growth of nonaxisymmetric perturbations, unless the sound
speed is at least a factor of two (Toomre's stability parameter )
higher that required by the usual Safronov-Toomre stability criteria
for axisymmetric perturbations. The strongest growth occurs on the scale of
.
In low and moderately high optical depth regions of the A and B rings,
cm s-1 and
g cm-2, thus
m, that is, the wavelength assumes a value of roughly several tens of particle diameters. These values of
and
m predicted in our analysis are close to Salo (1992, 1995), Richardson (1994), Osterbart & Willerding (1995), Griv (1998, 2005a), Daisaka & Ida (1999), Ohtsuki & Emori (2000), Griv & Gedalin (2005), and Griv et al. (2006a) numerical results. These fine waves were indeed found in data obtained by the C ASSINI spacecraft. The gravitation plays a key role in the formation of fine-scale structure.
Once a disk is violently unstable (the Safronov-Toomre criterion Q<1), both axisymmetric
and nonaxisymmetric
modes of comparable wavelength should grow at comparable rates and the
medium fragments into patches of particles with wavelength
in which the pressure cannot prevent the fragmentational collapse. The
geometric model of self-gravity spiral density waves in Saturn's rings
we study is shown in Fig. 2.
In the lowest approximation of the theory, the density waves are
regularly spaced, aligned three-dimensional clusters of ring particles
(Fig. 2a). At the limit of the instability (
), the physical scale of the waves in the equatorial z=0 plane is predicted to be of the order of
We conclude that the most unstable modes of the differentially rotating disk have wavelengths of some



![]() |
Figure 3:
Schematic model of a Jeans-unstable disk: a) the Safronov-Toomre unstable disk (
|
Open with DEXTER |
As for the present study, the Jeans-unstable spiral density waves also cause the ring A's quadrupole azimuthal brightness asymmetry detected first by Camichel (1958) and then by Lumme & Irvine (1976), Lumme et al. (1977), Thompson et al. (1981), Franklin et al. (1987), Dones et al. (1993), Dunn et al. (2004), and others (see Griv & Gedalin 2003; and Porco et al. 2008, for a discussion). The physics underlying these self-gravity waves is essentially the same as the ``density wave structure'' which was studied in the context of galactic disks by Lin & Shu (1966), Lin et al. (1969), Shu (1970), Lau & Bertin (1978), Lin & Lau (1979), Bertin (1980), Morozov (1980), Montenegro et al. (1999), Griv et al. (2002), and others.
![]() |
Figure 4:
a) Gravitationally unstable density waves with m=1 arm in the (
|
Open with DEXTER |
4.3 Formation of clumps
Thus, if the disk is thin,
,
and dynamically cold, that is,
(Toomre's stability parameter Q
< 1), then this model will be gravitationally unstable to both
axisymmetric and nonaxisymmetric perturbations, and should almost
instantaneously (see below for a time estimate) take on the form
of a cartwheel, that is, a structure of spirals and rings
(Fig. 3a). One understands,
however, that in the nonlinear stage of evolution there will be some
exchange of energy and angular momentum between the axisymmetric and
nonaxisymmetric modes that will give rise to a pattern far more complex
than the cartwheel shown in
Fig. 3a. The real spiral pattern is unlikely to be as regular as illustrated in Fig. 3a. It probably resembles the spiral pattern of ragged, multiarmed Sc galaxies (see Griv 2005b,
for a discussion). The ragged, ``irregular'' structure formation
process in planetary rings was demonstrated in the simulations carried
out by Salo (1992, 1995) and others (e.g., Daisaka & Ida 1999; Griv 2005a; Griv & Gedalin 2005) and observed in C ASSINI experiments (e.g., Colwell et al. 2007, p. 142, ``the sheets are loosely organized into the trailing spiral density enhancements''). C ASSINI
observations indicated that axisymmetric periodic microstructure in
Saturn's rings coexists with nonaxisymmetric structure (Colwell 2006,
2007; Thomson et al. 2007).
Clearly, in this case of both radial and spiral excitation, the
distribution of the surface density along the spiral arms is not
uniform, but describes a sequence of maxima, that might be identified
with forming embedded clusters of particles, resembling a ``beads on a
string'' structure. We argue that the gravitationally bound clusters of
particles appear at the intersections of the rings and spiral arms as
seen in Fig. 3a
. This disk should break up into discrete, gravitationally confined, and
porous blobs of matter (``clumpy moons'') of preferred mass
and diameters
distributed in spirals around the spin axis (Fig. 4). In Saturn's rings, the Safronov-Toomre sound speed

These values for the ring disk is based on a surface density of 50 g cm-2 and an angular speed of 2

We conclude that the Safronov-Toomre unstable modes might be
a potential cluster-forming mechanism (cf. Snytnikov et al. 2004, Figs. 2 and 4 therein; Griv 2005b, Fig. 2 therein). Since the intersections are spaced
apart, it would produce some regularity in the distance intervals
between the moon-clusters. Interestingly, low values of thermal
inertias of B and C ring particles derived from infrared
observations of Saturn's rings might be characteristic of very porous
particle aggregates (Ferrari et al. 2005). Tiscareno et al. (2006b, 2008, 2010) and
Srem
evi
et al. (2007) presented an extensive data set of localized features - ``propellers'' - from C ASSINI
images of Saturn's A ring that may be interpreted as signatures of
small moonlets embedded within the ring, with diameters between 40
and 500 m (cf. estimates (15) and (16)).
It was noted in particular that the lack of significant
brightening at high phase angle indicates that these bodies are likely
composed primarily of macroscopic particles, rather than dust
.
We strongly believe that the clumping Safronov-Toomre instability leads
to formation of porous aggregations of multiple smaller objects
(moonlets) with diameters
m embedded in broad Saturn's rings
.
4.4 Nonlinear evolution
The nonlinear interaction of particles with almost aperiodically
growing Jeans-unstable density waves increases the random velocities of
particles (via the so-called nonresonant dynamical heating) on a
short timescale of only 2-3 disk orbital revolutions (Griv &
Gedalin 2003; Griv et al. 2003a,b). The latter leads to an eventual stabilization (Eq. (A.34)),
unless some effective cooling mechanism of the reconstruction of the
wave structure exists. It is again suggested that dissipative
(inelastic) impacts between particles provide such a cooling mechanism,
reducing the magnitude of the relative velocity of particles, and thus
reducing the random velocity
spread. In particular we note that the particles constituting
Saturn's rings collide experiencing a subsequent loss of mechanical
energy. This mechanical energy may be converted into thermal energy,
and is essentially lost. The latter leads to some ``cooling'' of the
system, and thus leads to a recurrent instability cycle (Griv &
Gedalin 2003).
We argue that both the Saturnian main rings and their irregular
fine-scale structure (being recurrently created and destroyed on the
timescale of an order of Keplerian period h) are likely not much younger than the solar system (Griv & Gedalin 2006). Observations of the Saturnian ring system during approach and orbital insertion, with
C ASSINI's
visual and infrared mapping spectrometer, have already shown that
Saturn's rings have changed little in their radial structure since the
V OYAGER flybys in the early 1980s (Brown et al. 2006).
5 Discussion
For the first time in planetary ring dynamics, this paper has examined analytically the stability of three-dimensional Saturn's rings using a hydrodynamic approximation to particle dynamics. The linear analysis is restricted to odd-parity Jeans-type collective modes of oscillations - the Lin-Shu-type compression density waves - for which the vertical velocity vz(z) = - vz(-z) and, in particular vz(z=0)=0. It is assumed that Jeans' instability of small-amplitude gravity disturbances is the preferred mechanism to explain both the axisymmetric and the nonaxisymmetric structures present in the A and B rings (and probably in the C ring). The effect on the Saturnian ring disk structure by the dissipative effects (via interparticle collisions) seems likely to play a much smaller role. By assuming that equilibrium rotation and density vary over a much larger spatial scale than the mode wavelength, we found that the most important Jeans' gravitational instability in the Saturnian ring disk is a local instability, i.e., normal modes are driven unstable by the local value of the rotational flow shear and the disk gravity.
In previous sections, we have presented some arguments in favour of wave nature of the fine-scale m
quasi-periodic structure in Saturn's main rings, which is nothing else
but a free spiral density wave. From the well-developed theory of
galactic spiral density waves, a free density wave is known to
rotate in a rigid-body manner and to not be affected by differential
rotation of the ring disk.
Thus, one can attribute the fine-scale structure observed in Saturn's A and B rings in C ASSINI
data to the development of free Lin-Shu density waves developing in the
plane of the system. To repeat ourselves, these fully
self-consistent density waves, which are an intrinsic property of a
gravitationally unstable disk, are not to be confused with Julian &
Toomre's (1966) ``forced''
density wave proposal explored by Colombo et al. (1976), Franklin & Colombo (1978), Lumme & Irvine
(1979), and Karttunen (1983). In the spirit of Lin & Shu (1966), Lin et al. (1969), Shu (1970),
Ginzburg et al. (1972), Lau & Bertin (1978), Lin & Lau (1979), Bertin (1980), Morozov (1980),
Griv et al. (1999, 2002), Montenegro et al. (1999), Griv & Gedalin (2003), and others, we develop
instead the self-consistent theory of real instabilities of small spontaneous
disturbances, which grow and propagate effectively in the main part of
the system. In our model, the disk particles travel almost
freely through the density pattern, so any individual spiral arm
(and ring) is not always composed of the same particles; spirals rotate
rigidly with a single fixed pattern speed
,
hence do not wind up. In sharp contrast to the Julian-Toomre-type (Julian & Toomre 1966)
mechanism for producing the azimuthal (and radial) density variations,
the model advocated in this paper does not ``require the rings to
contain
particles whose radii are considerably larger than average'' (Colombo
et al. 1976). At the limit of
gravitational stability, the physical scale of these longitudinal waves (the width and separation of
the waves)
(Eq. (14))
and the disk's thickness is 2h. In Saturn's rings,
m. One concludes that the self-gravity density enhancements as shown in Fig. 2 are
flattened structures, with the width and the spacing between them of the order of ten particle
diameters and with height/width ratio of about
.
To reiterate, the self-gravity structures are flattened, having a vertical thickness only
of their radial extent; the inter-density enhancement spacing is
comparable to the width of the density enhancement itself.
A separate investigation based on high-resolution of the order of
10 m observations of Saturn's A and B (and
probably C) rings should be performed to confirm (or deny!)
this our prediction.
We have shown that the self-gravitating rings are able to form
various three-dimensional structures - ``sausage-like'' moderately
tightly-wound spirals and rings (Figs. 1b and 3), and clumpy moons (``rubble piles'') (Fig. 4),
continually forming and dispersing - without the interference of
external forces such as embedded or external satellites. Saturn's main
rings are likely a dynamic environment, in which the self-gravity
Lin-Shu density waves form and dissipate, and gaps and particle clumps
form and vanish. Gravitational instability of both ring and spiral
perturbations is found to provide an explanation of certain behaviors
observed by Tiscareno et al. (2006b, 2008) and Srem
evi
et al. (2007)
for the localized features of Saturn's A ring -
``propellers'' - which may be interpreted as signatures of
100-m-diameter porous moonlets of preferred mass
g each embedded within the broad ring. Although this has also yet to be directly measured.
The main shortcoming in our presentation is the short-wave
WKB assumption under which all the below considerations are only
valid. Without this assumption, Fourier normal modes are no longer
solutions of the linearized equations and one has to switch to far more
complicated mathematics (e.g., R
diger & Kitchatinov 2000).
We intend to address the issue in the following publications of the
series. Future work should also include detailed mechanisms of the
inelastic interaction: spin degrees of freedom, the particle size
distribution, and the finite size of the particles (e.g., Shukhman 1984; Araki 1991).
For instance, since the spin of a particle in Saturn's rings is
comparable to the orbital frequency, a large amount of energy may be
stored in this degree of freedom (e.g., Spilker et al. 2006; Leyrat et al. 2008b).
The latter increases the dissipativity of the system, and thus,
accelerates the formation of density waves. In local simulations
by Salo (1992, 1995)
and others, one can see spiral wave fragments (cylindric structures of
100 m or so) with a definite pitch angle with respect to the
local shear flow. This angle is almost about
.
In Saturn's rings, the preferred orientation of excited
gravitationally unstable density waves produces brightness azimuthal
variations similar to those observed by Camichel (1958), Lumme & Irvine (1976), Lumme et al. (1977), Thompson et al. (1981), Franklin et al. (1987), Dones et al. (1993), Dunn et al. (2004), French et al. (2007), and others. Even though the analysis presented here shows that there is a dominant nonaxisymmetric (
)
Fourier mode of maximum instability with
m, at the present time we cannot explain the angle
in the local WKB version of our theory. We also agree with the
referee of the paper that in terms of the production of clumpy moonlets
this work is highly undeveloped. At present, the paper would have
to be regarded as a prediction of moonlet formation by disk instability
based on theoretical modeling rather than as an explanation of observed
features. Future works should also consider large-scale simulations of
Saturn's rings. They could be used both to support the linear
dispersion analysis presented here and to examine the nonlinear growth
of the instability, especially in terms of the production of porous
moonlets.
The authors would like to thank Edward Liverts, Yury Lyubarsky, and Michael Mond for stimulating discussions and Irena Zlatopolsky for technical assistance and many comments, which increased the quality of the paper. We are indebted to Alexei M. Fridman for introducing this subject to one of us (E.G.) in 1986 and for his encouragement. We had the pleasure of collaborating with Tzi-Hong Chiueh and Ing-Guey Jiang. This work was sponsored by the Israel Science Foundation and the Binational US - Israel Science Foundation. The efforts of Evgeny Griv for this work were supported in part by the Israeli Ministry of Immigrant Absorption in the framework of the programme ``KAMEA''.
Appendix A: Stability of a three-dimensional disk
A.1 Three-dimensional perturbation
The time-dependent density
,
the gravitational potential
,
the pressure
P (r,t), and the fluid velocity
v (r,t) of a spatially inhomogeneous (only along the r coordinate) disk are represented by
where X(r,t) is any of the above-mentioned physical variables, X0(z) describes the basic flow, and






where






This is accurate for short wave perturbations only,

The latter releases us from taking into account the boundary conditions - one indeed considers the disk to be essentially infinite. The radial wavenumber

where




where now












In the local WKB approximation, it is assumed that the wave vector
and the wavefrequency vary continuously. By utilizing the more
accurate nonlocal approximation, it may be shown that the
characteristic oscillation frequencies of an inhomogeneous disk must
indeed be quantized, i.e., must pass through a discrete series of
values (Alexandrov et al. 1984,
p. 243). The existence of solutions to linearized gasdynamic
equations for a differentially rotating gravitating medium of the form
adopted here was examined by Lominadze et al. (1988) and Fridman (1989). Following Fridman (1989, p. 488), we assume that unstable perturbations develop rapidly on the instability timescale
(see Fridman 1989, for the definition of
). We refer the interested reader to Lominadze et al. (1988) and Fridman (1989) for details.
A.2 Solutions of the Poisson equation
Let us consider a disk of particles confined between two boundaries with their unperturbed coordinates, at z = +h and z = -h (Fig. 1a). A disk rotation with the angular velocity vector
parallel to the rotational axis z may be assumed. In this configuration, for this Lin-Shu-type perturbation the Poisson equation given by Eq. (4) takes a form
where


where



therefore, in these regions a solution may be sought that is similar to
where A, B, C, and D are to be determined and

![[*]](/icons/foot_motif.png)


(The potential



Solutions given by Eq. (A.12) is a new result of this work. By using a simplified analysis, very similar equations were already derived by Sekiya (1983, his Eqs. (3.6) and (3.7)), who examined the gravitational instability of the dust layer in the solar nebula. In sharp contrast to our study, however, Sekiya (1983) did not analyze the most important nonaxisymmetric perturbations propagating in a nonuniformly rotating and spatially inhomogeneous disk system.
One can use the first solution of Eqs. (A.12) to describe perturbations that are symmetric with respect to the z=0 equatorial plane of the disk (which do not cause it to bend) (Fig. 1b). For these even perturbations with
,
that is, A=D, from Eqs. (A.8) and (A.9) we have B =C. The system of of Eqs. (A.8)-(A.11) is then reduced to
from which the required relation, namely


After some trigonometric expansions and considering the most important cases of even long-wavelength oscillations (both |k|h<1 and qh<1), from the first solution given by Eq. (A.12) one
obtains
The short-wavelength perturbations, |k|h > 1, are not as dangerous in the problem of system stability as oscillations with |k|h < 1, since they lead only to very small-scale disturbances of the density with the radial scale






The second solution of Eq. (A.12),
,
describes odd perturbations
with
,
that is, where A=-D and B=-C. This type of vertical motion makes the disk bend in the same way as the plane of an oscillating membrane does (Fig. 1c). The vertical velocity of these motions is an even function of z:
vz(-z)=vz, and in the plane z=0,
it is not equal to zero. In contrast to the case of even
perturbations just considered above, the perturbed pressure, density,
gravitational potential, and horizontal velocity components are odd
functions of z. These gravity perturbations do not release
gravitational energy and, therefore, are expected to be
Jeans-stable (Bertin & Casertano 1982). The bending type of motions can be either caused by tidal influence of a satellite (Shu 1984), or excited by the so-called firehose-type bending instability (e.g., Griv & Chiueh 1998). Kulsrud et al. (1971) and Mark (1971)
investigated the bending instability developing in nonrotating disks by
using an energy principle. They demonstrated that the instability is
driven by the stellar ``pressure'' anisotropy: the source of free
energy in the instability is the intrinsic planar-to-vertical
anisotropy of a velocity dispersion. The firehose instability is well
known in plasma physics for transferring energy from one degree of
freedom to another (perpendicular) degree of freedom (e.g., Ichimaru 1973). Raha et al. (1991), Griv & Chiueh (1998), Liverts et al. (2003, Fig. 2 therein), Snytnikov et al. (2004, Fig. 5 therein), and Sotnikova & Rodionov (2005) presented nonresonant bending-unstable oscillations of three-dimensional rotating N-body models. The study of bending oscillations is beyond the scope of the present paper.
A.3 Dispersion relation
We next use the momentum equation given by Eq. (1)
to determine the perturbed velocity of the fluid element. Based on the
assumption that random velocities are anisotropic, the equations
of three-dimensional motion of the fluid element (Eq. (1)) in the frame of reference rotating with angular velocity
at the reference position r0 can be written in Hill's approximation as (Goldreich & Lynden-Bell 1965b)
where










In the absence of any perturbing gravity,

where K, N,





Equations (A.14)-(A.16) must be solved simultaneously with the solutions (A.12) and the continuity equation (see Eq. (2))
where








In Eqs. (A.21)-(A.23),
![\begin{displaymath}[ \cdots ]= \left( \omega_*^2 - \kappa^2 \right) \omega_* + \...
...th \eta k_{\rm r}^2 2 r \Omega \frac{{\rm d}\Omega}{{\rm d}r},
\end{displaymath}](/articles/aa/full_html/2010/13/aa14412-10/img242.png)


![]() |
|||
![]() |



The parameter









A special analysis of the solution close to corotation (
)
and Lindblad resonances is required. The Lindblad resonances occur when
the mean motion of the wave and that of the disk particles are in
the ratio
and in Saturn's rings



The set of Eqs. (A.20)-(A.23) is a system of algebraic equations. Using Eqs. (A.13) and (A.21)-(A.23), from Eqs. (A.4) and (A.20), it is straightforward to show that (e.g., Griv 2006)
where
is the square of the Jeans frequency and



Our dispersion relation given by Eq. (A.25) reduces to those of Lau & Bertin (1978), Lin & Lau (1979), Morozov (1985), Montenegro et al. (1999), and Griv et al. (2002)
when the finite thickness, ``out-of-phase'', and resonant contributions
vanish, so it seems as correct as their result. Equation (A.25) is complicated: it is highly nonlinear in the frequency .
To deal with the most interesting oscillation types analytically,
we consider various limiting cases of perturbations described by some
simplified variations in the dispersion relation. We first restrict
ourselves to considerating the principal part of the system between the
inner and outer Lindblad resonances,
,
and in particular, to the transparency region between the turning points (
)
in a disk. In the opposite case,
,
the effect of the disk rotation is negligible and therefore
irrelevant to us. Therefore, this limit never applies to the
rapidly rotating subsystems of both spiral galaxies and planetary
rings. Second, we can solve Eq. (A.25) by successive approximations. In the low-frequency (
)
and local WKB approximations (
)
that we indeed explore in Eq. (A.25), the terms that describe tangential forces
are assumed to be small compared to other terms and, therefore,
may be neglected to the lowest approximation. In addition,
the viscous effects are also considered to be weak,
.
A.4 Gravitational instability
From Eq. (A.25) for the most important high-frequency range in which we are interested
we determine the dispersion law for the Jeans branch of oscillations. By neglecting the small terms



(Lin et al. 1969). Accordingly, the disk is Jeans' gravitationally unstable (
)
to axisymmetric perturbations if
where
is the ordinary Safronov-Toomre (Safronov 1960, 1980; Toomre 1964) critical sound speed to suppress the instability of only axisymmetric (or radial) perturbations.
In the next approximation, in Eqs. (A.25) and (A.28), considering the Safronov-Toomre stable disk (
)
and the most important low-frequency
oscillations (
), in the small terms
one can
replace
by
.
As a result, the dispersion relation is obtained in the form
where
,
,
is now the square of the Jeans frequency,
is the squared effective wavenumber, and in Saturn's rings
.
Equation (A.30) differs from the ordinary Lin-Shu dispersion relation (A.28) by the appearance of the total k and effective k* wavenumbers,
which originate from the consideration of the nonaxisymmetrical modes
,
by the factor
,
which originates from the consideration of the effects of the finite thickness, and by the factor
,
which originates from the consideration of the viscous effects. Lynden-Bell & Kalnajs (1972, their Eq. (A11)) first obtained the Lin-Shu-type dispersion relation for open (
)
waves propagating in a homogeneous, dissipationless disk. A simplified dispersion relation (A.30) for low-frequency
perturbations that we are interested in can easily be obtain from Eqs. (D12) (in Eq. (D12) it should be
instead of T1; Montenegro et al. 1989) and (D14) of Lin & Lau (1979) by ignoring the viscous effects and the ``out-of-phase'' term
,
and using the expansion
.
From Eq. (A.30), we determine the dispersion law for the Jeans branch of oscillations
where p=1 for gravity-stable perturbations with
,
for gravity-unstable perturbations with
,
and the term involving
is the small correction (
).
Equation (A.32)
determines the spectrum of oscillations. Accordingly, viscosity leads
to a weak damping of both Jeans-unstable and Jeans-stable density
waves. Thus, even though the damping rate of the instability due to
collisions (
)
is not constant (as the density increases in the nonlinear part of the instability, e.g., Daisaka et al. 2001), this does not modify the papers' conclusions.
It follows from Eq. (A.32) that the growth rate of the instability is high,
,
and in general
that is, the instability develops rapidly on a dynamical timescale.
Equation (A.31) indicates the stabilizing influence of a finite thickness. Use of the dispersion curve minimum condition,
,
in Eq. (A.31) determines the wavelength of the most unstable Jeans mode
where




where







From Eqs. (A.29) and (A.34), it follows that once a disk is unstable according to the above criterions, namely
,
both axisymmetric modes and nonaxisymmetric modes of comparable
wavelengths, should increase at comparable rates; the disk should
break up in patches. It also follows that the viscosity has no
important effect on the criterion of the gravitational instability.
Equation (A.34) improves the Safronov-Toomre stability criterion
by including destabilizing effects resulting from shear
and azimuthal forces
(Griv et al. 2002; Griv 2006), and a stabilizing effect resulting from finite thickness
.
Maxwell (1859) considered just this kind of spiral instabilities with m=1
in his study concerning the stability of the Saturnian uniform rings
whose radial extent was considerably larger than the average
interparticle distance. Maxwell assumed that, in such a system,
the azimuthal force resulting from azimuthal displacements was more
important in determining the stability than the radial force produced
by radial displacements. In particular, Goldreich &
Lynden-Bell (1965a, p. 123) noted that ``
in the galaxy the tangential modes may be most unstable''. Lau & Bertin (1978, p. 509) (see also Bertin 1980; Morozov 1980)
then clarified the problem by considering the motion of a fluid
element: the density response that is in phase with the potential
minimum is found to exceed, by an amount proportional to both
and m, the corresponding response due to an axisymmetric field of equal strength. In Toomre (1981),
this amplification was discussed in terms of a ``swing mechanism'',
very reminiscent of the way we reach the modified stability
criterion (A.34). (The swing-amplified mechanism was discovered by Goldreich & Lynden-Bell 1965b.
The swing works on leading waves and turns them into trailing waves
producing strong amplification in the process.) The free kinetic
energy
associated with the differential rotation of the system is one possible
source for the growth of the energy of these spiral gravity
perturbations, and appears to be released when angular momentum is
transferred outward.
Vandervoort (1970), Yue (1982), Shu (1984), Morozov (1981), Romeo (1992, 1994), and Osterbart & Willerding (1995) already obtained somewhat similar stabilizing factors
by using simplified theories.
A.5 Dissipative instability
In contrast to Eq. (A.27), in the frequency range
Eq. (A.25) has another root equal to
where







As seen, the instability develops even in a rigidly rotating system. Because the limit




Following Lynden-Bell & Pringle (1974), Mishurov et al. (1976), and Morozov et al. (1985),
we also call the instability the dissipative instability because it is
introduced by collisions
between particles (and the disk's self-gravity). In our analysis,
we make the basic physical assumption that the transport
coefficients
and
are independent of state variables, in particular of the density
.
We have just shown above that these viscous-unstable perturbations grow aperiodically (
).
We note that the instability has nothing to do with the viscous
instability (viscous overstability) discussed by Borderies et al. (1985) and Schmit & Tscharnuter (1995).
In general, the effect of viscosity is to disrupt the organized wave
motion. A wave therefore tends to be damped on a timescale
of
(Bohm & Gross 1949). Lynden-Bell & Pringle (1974),
however, demonstrated that not all the organized motion will be lost in
a self-gravitating system: a dissipative type instability can
develop even in a Jeans-stable, rigidly
rotating self-gravitating viscous disk. According to Morozov
et al. (1985), the introduction of differential rotation leaves the result of Lynden-Bell & Pringle (1974) unchanged. Morozov et al. (1985)
also obtained the important result that the growth rate of the
dissipative instability reaches a maximum in the part of the disk that
is marginally Jeans-stable gravitationally. We refer to Willerding (1992) for the effect of simultaneous action of viscosity and self-gravity on rotating disks.
The cause of the instability, which has an essential dependence on the
self-gravitation of the disk matter, was explained in Lynden-Bell &
Pringle (1974), Mishurov et al. (1976), and Morozov et al. (1985).
Lynden-Bell & Pringle claimed this instability to be analogous to
the well-known viscous mechanism that converts Maclaurin spheroids to
Jacobi ellipsoids. The dissipative instability, as well as the
Jeans one studied above, can be considered as generating mechanisms for
unstable short-scale density waves, and these waves might be
responsible for the appearance of structures in galaxies, Saturn's
rings, and protoplanetary disks (Morozov et al. 1985; Gorkavyi & Fridman 1990a,b). In the framework of our theory, however, the Saturnian ring disk is gravitationally unstable (
)
and therefore the weak (
)
dissipative instability in Saturn's rings studied in this section is
unlikely.
We find evidence that the growth rate of the dissipative instability is
indeed small in the case of low and relatively high optical depth rings
making Jeans' gravitational instability studied in Sect. A.4
above a leading candidate mechanism to explain both axisymmetric and
nonaxisymmetric microstructures observed in Saturn's A and
B rings (and probably in its C ring as well).
References
- Alexander, R. D., Armitage, P. J., Cuadra, J., & Begelman, M. C. 2008, ApJ, 674, 927 [NASA ADS] [CrossRef] [Google Scholar]
- Alexandrov, A. F., Bogdankevich, L. S., & Rukhadze, A. A. 1984, Principles of Plasma Electrodynamics (New York: Springer) [Google Scholar]
- Altobelli, N., Spilker, L. J., Leyrat, C., & Pilorz, S. 2008, Planet. Space Sci., 56, 134 [NASA ADS] [CrossRef] [Google Scholar]
- Araki, S. 1991, Am. Sci., 79, 44 [NASA ADS] [Google Scholar]
- Arsenin, V. V. 1967, Soviet Phys. - Tech. Phys., 12, 442 [Google Scholar]
- Bertin, G. 1980, Phys. Rep., 61, 1 [NASA ADS] [CrossRef] [Google Scholar]
- Bertin, G. 2000, Dynamics of Galaxies (Cambridge: Cambridge Univ. Press) [Google Scholar]
- Bertin, G., & Casertano, S. 1982, A&A, 106, 274 [NASA ADS] [Google Scholar]
- Bertin, G., & Lin, C. C. 1996, Spiral Structure in Galaxies: A Density Wave Theory (Cambridge: MIT Press) [Google Scholar]
- Bohm, D., & Gross, E. P. 1949, Phys. Rev., 75, 1864 [NASA ADS] [CrossRef] [Google Scholar]
- Borderies, N., & Longaretti, P.-Y. 1994, Icarus, 107, 129 [NASA ADS] [CrossRef] [Google Scholar]
- Borderies, N., Goldreich, P., & Tremaine, S. 1985, Icarus, 63, 406 [NASA ADS] [CrossRef] [Google Scholar]
- Boss, A. 2005, ApJ, 629, 535 [NASA ADS] [CrossRef] [Google Scholar]
- Boss, A. 2007, ApJ, 661, L73 [NASA ADS] [CrossRef] [Google Scholar]
- Boss, A. 2008, ApJ, 677, 607 [NASA ADS] [CrossRef] [Google Scholar]
-
Brahic, A. 2001, in Granular Gases, ed. T. P
schel, & A. Luding (Berlin: Springer), 281 [Google Scholar]
- Bridges, F. G., Hatzes, A. P., & Lin, D. N. C. 1984, Nature, 309, 333 [NASA ADS] [CrossRef] [Google Scholar]
- Brown, R. H., Baines, K. H., Bellucci, G., et al. 2006, A&A, 446, 707 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Camichel, H. 1958, Ann. d'Astr., 21, 231 [NASA ADS] [Google Scholar]
- Chagelishvili, G. D., Zahn, J.-P., Tevzadze, A. G., & Lominadze, J. G. 2003, A&A, 402, 401 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Chandrasekhar, S. 1955, Vistas Astron., 1, 344 [NASA ADS] [CrossRef] [Google Scholar]
- Clampin, M., Krist, J. E., Ardila, D. R., et al. 2003, AJ, 126, 385 [NASA ADS] [CrossRef] [Google Scholar]
- Colombo, G., Goldreich, P., & Harris, A. 1976, Nature, 264, 344 [NASA ADS] [CrossRef] [Google Scholar]
-
Colwell, J. E., Esposito, L. W., & Srem
evi
, M. 2006, Geophys. Res. Lett., 33, L07201 [NASA ADS] [CrossRef] [Google Scholar]
-
Colwell, J. E., Esposito, L. W., Srem
evi
, M., Stewart, G. R., & McClintock, W. E. 2007, Icarus, 90, 127 [NASA ADS] [CrossRef] [Google Scholar]
-
Colwell, J. E., Cooney, J. H., Esposito, L. W., & Srem
evi
, M. 2009, Icarus, 200, 574 [NASA ADS] [CrossRef] [Google Scholar]
- Cook, A. F., & Franklin, F. A. 1964, AJ, 69, 17 [Google Scholar]
- Cossins, P., Lodato, G., & Clarke, C. J. 2009, MNRAS, 393, 1157 [Google Scholar]
- Cuzzi, J. N., Colwell, J. E., Esposito, L. W., et al. 2002, Space Sci. Rev., 118, 209 [NASA ADS] [CrossRef] [Google Scholar]
- Cuzzi, J., Clark, R., Filacchione, G., et al. 2009, in Saturn from Cassini-Huygens, ed. M. K. Dougherty, L. W. Esposito, & S. M. Krimigis (Netherlands: Springer), 459 [Google Scholar]
- Daisaka, H., & Ida, S. 1999, Earth Planets Space, 51, 1195 [NASA ADS] [Google Scholar]
- Daisaka, H., Tanaka, H., & Ida, S. 2001, Icarus, 154, 296 [NASA ADS] [CrossRef] [Google Scholar]
- Dodson-Robinson, S. E., Veras, D., Ford, E. B., & Beichman, C. A. 2009, ApJ, 707, 79 [NASA ADS] [CrossRef] [Google Scholar]
- Dones, L., Cuzzi, J. N., & Showalter, M. R. 1993, Icarus, 105, 184 [NASA ADS] [CrossRef] [Google Scholar]
- Dunn, D. E., Molnar, L. A., Niehof, J. T., de Pater, I., & Lissauer, J. J. 2004, Icarus, 171, 183 [NASA ADS] [CrossRef] [Google Scholar]
- Durisen, R. H., Boss, A. P., Mayer, L., et al. 2007, in Protostars and Planets V, ed. B. Reipurth, D. Jewitt, & K. Keil (Tucson: Univ. Arizona Press), 607 [Google Scholar]
- Esposito, L. W. 2002, Rep. Prog. Phys., 65, 1741 [Google Scholar]
- Esposito, L. W., O'Callaghan, M., & West, R. A. 1983, Icarus, 56, 439 [NASA ADS] [CrossRef] [Google Scholar]
- Ferrari, C., Galdemard, P., Lagage, P. O., Pantin, E., & Quoirin, C. 2005, A&A, 441, 379 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Ferrari, C., Brooks, S., Edgington, S., et al. 2009, Icarus, 199, 145 [NASA ADS] [CrossRef] [Google Scholar]
- Franklin, F. A., & Colombo, G. 1978, Icarus, 33, 279 [NASA ADS] [CrossRef] [Google Scholar]
- Franklin, F. A., Cook, A. F., Barrey, R. T. F., et al. 1987, Icarus, 69, 280 [NASA ADS] [CrossRef] [Google Scholar]
- French, R. G., & Nicholson, P. D. 2000, Icarus, 145, 502 [NASA ADS] [CrossRef] [Google Scholar]
- French, R. G., Salo, H., McGhee, C. A., & Dones, L. 2007, Icarus, 189, 493 [NASA ADS] [CrossRef] [Google Scholar]
- Fridman, A. M. 1989, Soviet Ast. Lett., 15, 487 [Google Scholar]
- Fukagawa, M., Hayashi, M., Tamura, M., et al. 2004, ApJ, 605, L53 [NASA ADS] [CrossRef] [Google Scholar]
- Fukagawa, M., Tamura, M., Itoh, Y., et al. 2006, ApJ, 636, L153 [NASA ADS] [CrossRef] [Google Scholar]
- Galeev, A. A., & Sagdeev, R. Z. 1983, in Basic Plasma Physics, ed. M. N. Rosenbluth, & R. Z. Sagdeev (Amsterdam: North-Holland), 1, 679 [Google Scholar]
- Gammie, C. F. 2001, ApJ, 553, 174 [NASA ADS] [CrossRef] [Google Scholar]
- Ginzburg, I. F., Polyachenko, V. L., & Fridman, A. M. 1972, Soviet Ast., 15, 643 [NASA ADS] [Google Scholar]
- Goldreich, P., & Lynden-Bell, D. 1965a, MNRAS, 130, 97 [NASA ADS] [CrossRef] [Google Scholar]
- Goldreich, P., & Lynden-Bell, D. 1965b, MNRAS, 130, 125 [NASA ADS] [CrossRef] [Google Scholar]
- Goldreich, P., & Tremaine, S. 1982, ARA&A, 20, 249 [NASA ADS] [CrossRef] [Google Scholar]
- Gorkavyi, N. N., & Fridman, A. M. 1990a, Soviet Phys. - Uspehi, 33, 95 [NASA ADS] [CrossRef] [Google Scholar]
- Gorkavyi, N. N., & Fridman, A. M. 1990b, Soviet Ast. Lett., 16, 79 [NASA ADS] [Google Scholar]
- Griv, E. 1998, Planet. Space Sci., 46, 615 [Google Scholar]
-
Griv, E. 2005a, in Dynamics of Populations of Planetary Systems, ed. Z. Kne
evi
, & A. Milani (Cambridge: Cambridge Univ. Press), IAU Colloq., 197, 459 [Google Scholar]
- Griv, E. 2005b, Ap&SS, 299, 371 [NASA ADS] [CrossRef] [Google Scholar]
- Griv, E. 2006, A&A, 449, 573 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Griv, E. 2007a, ApJ, 665, 866 [NASA ADS] [CrossRef] [Google Scholar]
- Griv, E. 2007b, Planet. Space Sci., 55, 547 [NASA ADS] [CrossRef] [Google Scholar]
- Griv, E., & Chiueh, T. 1998, ApJ, 503, 186 [NASA ADS] [CrossRef] [Google Scholar]
- Griv, E., & Gedalin, M. 2003, Planet. Space Sci., 51, 899 [NASA ADS] [CrossRef] [Google Scholar]
- Griv, E., & Gedalin, M. 2004, AJ, 128, 1956 [Google Scholar]
- Griv, E., & Gedalin, M. 2005, Planet. Space Sci., 53, 461 [NASA ADS] [CrossRef] [Google Scholar]
- Griv, E., & Gedalin, M. 2006, Planet. Space Sci., 54, 794 [NASA ADS] [CrossRef] [Google Scholar]
- Griv, E., Rosenstein, B., Gedalin, M., & Eichler, D. 1999, A&A, 347, 821 [NASA ADS] [Google Scholar]
- Griv, E., Gedalin, M., Eichler, D., & Yuan, C. 2000, Planet. Space Sci., 48, 679 [Google Scholar]
- Griv, E., Gedalin, M., & Yuan, C. 2002, A&A, 383, 338 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Griv, E., Gedalin, M., & Yuan, C. 2003a, A&A, 400, 375 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Griv, E., Gedalin, M., & Yuan, C. 2003b, MNRAS, 342, 1102 [NASA ADS] [CrossRef] [Google Scholar]
- Griv, E., Gedalin, M., & Lyubarsky, Yu. 2006a, Adv. Space Res., 38, 770 [NASA ADS] [CrossRef] [Google Scholar]
- Griv, E., Gedalin, M., & Yuan, C. 2006b, Adv. Space Res., 38, 47 [NASA ADS] [CrossRef] [Google Scholar]
- Griv, E., Liverts, E., & Mond, M. 2008, A&A, 672, L127 [NASA ADS] [Google Scholar]
- Harris, E. G. 1968, in Physics of Hot Plasmas, ed. B. J. Rye, & J. C. Taylor (New York: Plenum), 145 [Google Scholar]
- Hedman, M. M., Nicholson, P. D., Salo, H., et al. 2007, AJ, 133, 2624 [NASA ADS] [CrossRef] [Google Scholar]
- Hedman, M. M., Burns, J. A., Tiscareno, M. S., & Porco, C. C. 2009, Planet. Space Sci., 202, 260 [Google Scholar]
- Hohl, F. 1972, J. Comput. Phys., 9, 10 [NASA ADS] [CrossRef] [Google Scholar]
- Hohl, F. 1978, AJ, 83, 768 [NASA ADS] [CrossRef] [Google Scholar]
- Holberg, J. B., Forrester, W. T., & Lissauer, J. J. 1982, Nature, 297, 115 [NASA ADS] [CrossRef] [Google Scholar]
- Horne, L. J., & Cuzzi, J. N. 1996, Icarus, 119, 285 [NASA ADS] [CrossRef] [Google Scholar]
- Ichimaru, S. 1973, Basic Principles of Plasma Physics (Reading: Benjamin) [Google Scholar]
- Jeans, J. H. 1928, Astronomy and Cosmogony (Cambridge: Cambridge Univ. Press) [Google Scholar]
- Jeffreys, H. 1947, MNRAS, 107, 263 [NASA ADS] [Google Scholar]
- Julian, W. H., & Toomre, A. 1966, ApJ, 146, 810 [NASA ADS] [CrossRef] [Google Scholar]
- Karjalainen, R., & Salo, H. 2004, Icarus, 172, 328 [NASA ADS] [CrossRef] [Google Scholar]
- Karttunen, H. 1983, Moon Planets, 29, 117 [NASA ADS] [CrossRef] [Google Scholar]
- Kerr, R. A. 1985, Science, 220, 1376 [NASA ADS] [CrossRef] [Google Scholar]
- Krall, N. A., & Trivelpiece, A. W. 1986, Principles of Plasma Physics (San Francisco: San Francisco Press) [Google Scholar]
- Kulsrud, R. M. 1972, in Gravitational N-body Problem, ed. M. Lecar (Dordrecht: Reidel), 337 [Google Scholar]
- Kulsrud, R. M., Mark, J. W.-K., & Caruso, A. 1971, Ap&SS, 14, 52 [NASA ADS] [CrossRef] [Google Scholar]
- Landau, L. D., & Lifshitz, E. M. 1987, Fluid Mechanics (Oxford: Buttenworth-Heinemann) [Google Scholar]
- Lane, A. L., Hord, C. W., West, R. A., et al. 1982, Science, 215, 537 [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Lau, Y. Y., & Bertin, G. 1978, ApJ, 226, 508 [NASA ADS] [CrossRef] [Google Scholar]
- Laughlin, G., & Bodenheimer, P. 1994, ApJ, 436, 335 [NASA ADS] [CrossRef] [Google Scholar]
-
Laughlin, G., & R
zyczka, M. 1996, ApJ, 456, 279 [NASA ADS] [CrossRef] [Google Scholar]
- Leyrat, C., Spilker, L. J., Altobelli, N., Pilorz, S., & Ferrari, C. 2008a, Planet. Space Sci., 56, 117 [NASA ADS] [CrossRef] [Google Scholar]
- Leyrat, C., Ferrari, C., Charnoz, S., et al. 2008b, Icarus, 196, 625 [NASA ADS] [CrossRef] [Google Scholar]
- Lin, C. C., & Shu, F. H. 1966, Proc. Natl. Acad. Sci., 55, 229 [Google Scholar]
- Lin, C. C., & Shu, F. H. 1968, in Astrophysics and General Relativity, ed. M. Chretian, S. Deser, & J. Goldstein (New York: Gordon and Breach), 2, 236 [Google Scholar]
- Lin, C. C., & Lau, Y. Y. 1979, SIAM Stud. Appl. Math., 60, 97 [NASA ADS] [CrossRef] [Google Scholar]
- Lin, D. N. C., & Paploizou, J. C. B. 1996, ARA&A, 34, 703 [NASA ADS] [CrossRef] [Google Scholar]
- Lin, C. C., Yuan, C., & Shu, F. H. 1969, ApJ, 155, 721 (erratum, 156, 797) [NASA ADS] [CrossRef] [Google Scholar]
- Lissauer, J. J., & Cuzzi, J. N. 1982, Icarus, 87, 1051 [Google Scholar]
- Liverts, E., Griv, E., Gedalin, M., & Eichler, D. 2003, in Galaxies and Chaos, ed. G. Contopoulos, & N. Voglis (Berlin: Springer), 340 [Google Scholar]
- Lominadze, D. G., Chagelishvili, G. D., & Chanishvili, R. G. 1988, Soviet Ast. Lett., 14, 364 [Google Scholar]
- Lumme, K., & Irvine, W. M. 1976, ApJ, 204, L55 [NASA ADS] [CrossRef] [Google Scholar]
- Lumme, K., & Irvine, W. M. 1979, Nature, 282, 695 [NASA ADS] [CrossRef] [Google Scholar]
- Lumme, K., Esposito, L. W., Irvine, W. M., & Baum, W. A. 1977, ApJ, 216, L123 [NASA ADS] [CrossRef] [Google Scholar]
- Lynden-Bell, D., & Kalnajs, A. J. 1972, MNRAS, 157, 1 [NASA ADS] [CrossRef] [Google Scholar]
- Lynden-Bell, D., & Pringle, J. E. 1974, MNRAS, 168, 603 [NASA ADS] [CrossRef] [Google Scholar]
- Maoz, E. 1995, ApJ, 455, L131 [NASA ADS] [CrossRef] [Google Scholar]
- Mark, J. W.-K. 1971, ApJ, 169, 455 [NASA ADS] [CrossRef] [Google Scholar]
- Mayer, L., Quinn, T., Wadsley, J., & Stadel, J. 2002, Science, 298, 1756 [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Mayer, L., Lufkin, G., Quinn, T., & Wadsley, J. 2007, ApJ, 661, L77 [NASA ADS] [CrossRef] [Google Scholar]
- Maxwell, J. C. 1859, in Sci. Papers, 1960, ed. W. D. Niven (New York: Dover), 1, 28 [Google Scholar]
- Mishurov, Yu. N., Peftiev, V. N., & Suchkov, A. A. 1976, Soviet Ast., 20, 152 [NASA ADS] [Google Scholar]
- Montenegro, L. E., Yuan, C., & Elmegreen, B. G. 1999, ApJ, 520, 592 [NASA ADS] [CrossRef] [Google Scholar]
- Morozov, A. G. 1980, Soviet Ast., 24, 391 [Google Scholar]
- Morozov, A. G. 1981, Soviet Ast. Lett., 7, 109 [Google Scholar]
- Morozov, A. G. 1985, Soviet Ast., 29, 120 [NASA ADS] [Google Scholar]
- Morozov, A. G., Torgashin, Yu. M., & Fridman, A. M. 1985, Soviet Ast. Lett., 11, 94 [NASA ADS] [Google Scholar]
- Mosqueira, I. 1996, Icarus, 122, 128 [NASA ADS] [CrossRef] [Google Scholar]
- Murray, C. D., Beurle, K., Cooper, N. J., et al. 2008, Nature, 453, 739 [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Nicholson, P. D., & Dones, L. 1991, Rev. Geophys. Suppl., 29, 313 [NASA ADS] [Google Scholar]
- Nicholson, P. D., & Hedman, M. M. 2010, Icarus, 206, 410 [NASA ADS] [CrossRef] [Google Scholar]
- Nicholson, P. D., French, R. G., Tollestrup, E., et al. 2000, Icarus, 145, 474 [NASA ADS] [CrossRef] [Google Scholar]
- Nicholson, P. D., French, R. G., Campbell, D. B., et al. 2005, Icarus, 177, 32 [NASA ADS] [CrossRef] [Google Scholar]
- Nicholson, P. D., Hedman, M. M., Clark, R. N., et al. 2008, Icarus, 193, 182 [NASA ADS] [CrossRef] [Google Scholar]
- Ohtsuki, K., & Emori, H. 2000, AJ, 119, 403 [NASA ADS] [CrossRef] [Google Scholar]
- Osterbart, R., & Willerding, E. 1995, Planet. Space Sci., 43, 289 [Google Scholar]
- Papaloizou, J. C. B., & Lin, D. N. C. 1995, ARA&A, 33, 505 [NASA ADS] [CrossRef] [Google Scholar]
- Porco, C. C. 1990, Adv. Space Res., 10, 221 [NASA ADS] [CrossRef] [Google Scholar]
- Porco, C. C., Baker, E., Barbara, J., et al. 2005, Science, 307, 1226 [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Porco, C. C., Weiss, J. W., Richardson, D. C., et al. 2008, AJ, 136, 2172 [NASA ADS] [CrossRef] [Google Scholar]
- Postnikov, E. B., & Loskutov, A. Yu. 2007, J. Exper. Theor. Phys., 104, 417 [NASA ADS] [CrossRef] [Google Scholar]
-
Pickett, B. K., Mej
a, A. C., Durisen, R. H., et al. 2003, ApJ, 590, 1060 [NASA ADS] [CrossRef] [Google Scholar]
- Raha, N., Sellwood, J. A., James, R. A., & Kahn, F. D. 1991, Nature, 352, 411 [NASA ADS] [CrossRef] [Google Scholar]
- Rice, W. K. M., Lodato, G., & Armitage, P. J. 2005, MNRAS, 364, L56 [NASA ADS] [CrossRef] [Google Scholar]
- Richardson, D. C. 1994, MNRAS, 269, 493 [NASA ADS] [CrossRef] [Google Scholar]
- Romeo, A. B. 1992, MNRAS, 256, 307 [NASA ADS] [CrossRef] [Google Scholar]
- Romeo, A. B. 1994, A&A, 286, 799 [NASA ADS] [Google Scholar]
- Rosen, P. A., Tyler, G. L., Marouf, E. A., & Lissauer, J. J. 1991, Icarus, 93, 25 [NASA ADS] [CrossRef] [Google Scholar]
-
R
diger, G., & Kitchatinov, L. L. 2000, Astron. Nachr., 21, 181 [Google Scholar]
- Safronov, V. S. 1960, Ann. Astrn., 23, 979 [Google Scholar]
- Safronov, V. S. 1980, in Early Solar System Processes, ed. D. Lal (Amsterdam: North-Holland), 73 [Google Scholar]
- Sagdeev, R. Z., & Galeev, A. A. 1969, Nonlinear Plasma Theory, ed. T. M. O'Neil, & D. L. Book (New York: Benjamin) [Google Scholar]
- Salo, H. 1992, Nature, 359, 619 [NASA ADS] [CrossRef] [Google Scholar]
- Salo, H. 1995, Icarus, 117, 287 [NASA ADS] [CrossRef] [Google Scholar]
- Schmit, U., & Tscharnuter, W. M. 1995, Icarus, 115, 304 [NASA ADS] [CrossRef] [Google Scholar]
- Sekiya, M. 1983, Prog. Theor. Phys., 69, 1116 [NASA ADS] [CrossRef] [Google Scholar]
- Sellwood, J. A., & Athanassoula, E. 1986, MNRAS, 221, 195 [NASA ADS] [Google Scholar]
- Sellwood, J. A., & Kahn, F. D. 1991, MNRAS, 250, 278 [NASA ADS] [Google Scholar]
- Sellwood, J. A., & Lin, D. C. 1989, MNRAS, 240, 991 [NASA ADS] [Google Scholar]
- Showalter, M. R., & Nicholson, P. D. 1990, Icarus, 87, 285 [NASA ADS] [CrossRef] [Google Scholar]
- Shu, F. H. 1970, ApJ, 160, 99 [Google Scholar]
- Shu, F. H. 1984, in Planetary Rings, ed. R. Greenberg, & A. Brahic (Tucson: Univ. Arizona Press), 513 [Google Scholar]
- Shu, F. H., & Stewart, G. R. 1985, Icarus, 62, 360 [NASA ADS] [CrossRef] [Google Scholar]
- Shu, F. H., Cuzzi, J. N., & Lissauer, J. J. 1983, Icarus, 53, 185 [NASA ADS] [CrossRef] [Google Scholar]
- Shu, F. H., Dones, L., Lissauer, J. J., Yuan, C., & Cuzzi, J. N. 1985, ApJ, 299, 542 [NASA ADS] [CrossRef] [Google Scholar]
- Shukhman, I. G. 1984, Soviet Ast., 28, 574 [Google Scholar]
- Sicardy, B. 2005, Space Sci. Rev., 116, 457 [NASA ADS] [CrossRef] [Google Scholar]
- Smith, B. A., Soderblom, L., Beebe, R. F., et al. 1981, Science, 212, 163 [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Smith, B. A., Soderblom, R., Batson, R., et al. 1982, Science, 215, 504 [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Snytnikov, V. N., Vshivkov, V. A., Kuksheva, E. A., et al. 2004, Astr. Lett., 30, 124 [NASA ADS] [CrossRef] [Google Scholar]
- Sotnikova, N. Ya., & Rodionov, S. A. 2005, Astr. Lett., 31, 15 [Google Scholar]
- Spilker, L. J., Pilorz, S. H., Wallis, B. D., et al. 2006, Planet. Space Sci., 54, 1167 [NASA ADS] [CrossRef] [Google Scholar]
-
Srem
evi
, M., Schmidt, J., Salo, H., et al. 2007, Nature, 449, 1019 [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Stewart, G., Lin, D. N. C., & Bodenheimer, P. 1984, in Planetary Rings, ed. R. Greenberg, & A. Brahic (Tucson: Univ. Arizona Press), 447 [Google Scholar]
- Stone, E. C., & Miner, E. D. 1982, Science, 215, 499 [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Swanson, D. G. 1989, Plasma Waves (Boston: Academic Press) [Google Scholar]
- Takeda, T., & Ida, S. 2001, ApJ, 560, 514 [NASA ADS] [CrossRef] [Google Scholar]
- Thompson, W. T., Lumme, K., Irvine, W. M., Baum, W. A., & Esposito, L. W. 1981, Icarus, 46, 187 [NASA ADS] [CrossRef] [Google Scholar]
- Thomson, F. S., Marouf, E. A., Tyler, G. L., French, R. G., & Rappoport, N. J. 2007, Geophys. Res. Lett., 34, L24203 [NASA ADS] [CrossRef] [Google Scholar]
- Tiscareno, M. S., Nicholson, P. D., Burns, J. A., Hedman, M. M., & Porco, C. C. 2006a, ApJ, 651, L65 [NASA ADS] [CrossRef] [Google Scholar]
- Tiscareno, M. S., Burns, J. A., Hedman, M. M., et al. 2006b, Nature, 440, 648 [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Tiscareno, M. S., Burns, J. A., Nicholson, P. D., Hedman, M. M., & Porco, C. C. 2007, Icarus, 189, 14 [NASA ADS] [CrossRef] [Google Scholar]
- Tiscareno, M. S., Burns, J. A., Hedman, M. M., & Porco, C. C. 2008, AJ, 135, 1083 [Google Scholar]
- Tiscareno, M. S., Perrine, R. P., Richardson, D. C., et al. 2010, AJ, 139, 492 [NASA ADS] [CrossRef] [Google Scholar]
- Tomley, L., Cassen, P., & Steiman-Cameron, T. 1991, ApJ, 382, 530 [NASA ADS] [CrossRef] [Google Scholar]
- Tomley, L., Steiman-Cameron, T., & Cassen, P. 1994, ApJ, 422, 850 [NASA ADS] [CrossRef] [Google Scholar]
- Toomre, A. 1964, ApJ, 139, 1217 [NASA ADS] [CrossRef] [Google Scholar]
- Toomre, A. 1981, in The Structure and Evolution of Normal Galaxies, ed. S. F. Fall, & D. Lynden-Bell (Cambridge: Cambridge Univ. Press), 111 [Google Scholar]
- Vandervoort, P. O. 1970, ApJ, 161, 87 [NASA ADS] [CrossRef] [Google Scholar]
- Willerding, E. 1992, Earth Moon Planets, 56, 173 [NASA ADS] [CrossRef] [Google Scholar]
- Yuan, C. 1969, ApJ, 158, 871 [NASA ADS] [CrossRef] [Google Scholar]
- Yuan, C. 2002, Space Sci. Rev., 102, 121 [NASA ADS] [CrossRef] [Google Scholar]
- Yue, Z. Y. 1982, Geophys. Astrophys. Fluid Dyn., 20, 1 [NASA ADS] [CrossRef] [Google Scholar]
- Zebker, H. A., Marouf, E. A., & Tyler, G. L. 1985, Icarus, 84, 531 [NASA ADS] [CrossRef] [Google Scholar]
Footnotes
- ... discussion)
- We consider the latter name, namely ``self-gravity wakes'', less than apt and will avoid it here (see also Colwell et al. 2006, for a discussion).
- ... spirals
- Destabilizing self-gravity in far more ``dangerous'' in thin
disks than in thick disks. If a rotating disk has a large vertical
thickness owing to a high internal temperature, then it is stabilized
against all gravitational instabilities. Instabilities arise as the
thickness of the layer is reduced (Safronov 1980; Shu 1984). We are interested only in thin astrophysical systems with a ratio of the half-thickness h to the outer radius R much smaller than unity:
. This ratio is characteristic of all spiral galaxies, the gaseous disks around black holes, the planet-forming disks of protostars, and planetary rings. The main Saturn's rings are only about 10 m thick.
- ... times
- If the nonaxisymmetric waves only consisted of the same particles, than these waves would ``wrap up'' around the planet and essentially smear out of visibility.
- ... galaxies
- Whether these Julian-Toomre-type forced density waves can be excited in astrophysical disks remains controversial (e.g., Sellwood & Lin 1989, p. 992; Sellwood & Kahn 1991, p. 278).
- ...
rings
- Salo (1992, 1995) and others (e.g., Griv 2005a) already demonstrated the formation of fine-scale structures in simplified local N-body simulations of planetary rings of identical particles.
- ... oscillations
- The disk is geometrically thin if
, which from the equation
, where
and
, is equivalent to the disk being dynamically cold,
, i.e., the sound speed of the ring disk being much less than the orbital speed
.
- ...a
- The latter is an important step towards an understanding of a main question of protoplanetary disk evolution, as well as the evolutionary processes in galactic disks: what kind of evolutionary processes lead to the formation of moons, planets, and stars in a different astrophysical disk system? Maoz (1995) already explained the apparent clustering of the high-velocity emission sources into several distinct clumps near the center of the galaxy NGC 4258 at the intersections of the spiral arms and the rings.
- ... dust
- A close-up of Saturn's rings reveals many bright streaks aligned with the orbital direction of the rings. These objects are the
propeller-shaped features first captured in C ASSINI
images during the spacecraft's 2004 orbital insertion manoeuvre.
Propeller-shaped disturbances occur in Saturn's rings as a result of
small moonlets embedded in the rings (Tiscareno et al. 2006). The
propellers are most abundant in a 3000 km-wide belt in the
mid-A ring, about 130 000 km from Saturn's center (Srem
evi
et al. 2007). It is estimated that the A ring contains thousands of these objects. Tiscareno et al. (2008) presented a detailed analysis of this population.
- ... rings
- Note that in the context of protoplanetary disks Toomre's stability parameter Q < 1 is sufficient to create point masses (giant planets) by disk instability (Boss 2005, 2007).
- ... number
- This is because for these longitudinal perturbations the equations in
Sect. A.3 below are linear in z (and r,
, t) and, therefore, we may Fourier-analyze them in the form
.
All Figures
![]() |
Figure 1: Sketch of perturbations of a three-dimensional disk. In a) a section of the disk is shown edge-on. In b) a mode of even symmetry with respect to the equatorial plane, or an even-parity Jeans-type perturbation, is shown (the dashed line). In c) a mode of odd symmetry with respect to the equatorial plane, or an odd bending-type perturbation, is illustrated (the dashed line). |
Open with DEXTER | |
In the text |
![]() |
Figure 2: Schematic model of the fine-scale density wave structure in Saturn's rings. Self-gravity density waves, which were first studied by Lin & Shu (Lin et al. 1969; Shu 1970), manifest themselves as evenly spaced elongated clusters of ring particles. Shown are both a three-dimensional distribution of particles a) and a distribution of particles in the (x,z)-plane b). |
Open with DEXTER | |
In the text |
![]() |
Figure 3:
Schematic model of a Jeans-unstable disk: a) the Safronov-Toomre unstable disk (
|
Open with DEXTER | |
In the text |
![]() |
Figure 4:
a) Gravitationally unstable density waves with m=1 arm in the (
|
Open with DEXTER | |
In the text |
Copyright ESO 2010
Current usage metrics show cumulative count of Article Views (full-text article views including HTML views, PDF and ePub downloads, according to the available data) and Abstracts Views on Vision4Press platform.
Data correspond to usage on the plateform after 2015. The current usage metrics is available 48-96 hours after online publication and is updated daily on week days.
Initial download of the metrics may take a while.