Articles

SPIRAL INSTABILITIES IN N-BODY SIMULATIONS. I. EMERGENCE FROM NOISE

Published 2012 May 3 © 2012. The American Astronomical Society. All rights reserved.
, , Citation J. A. Sellwood 2012 ApJ 751 44 DOI 10.1088/0004-637X/751/1/44

0004-637X/751/1/44

ABSTRACT

The origin of spiral patterns in galaxies is still not fully understood. Similar features also develop readily in N-body simulations of isolated cool, collisionless disks, yet even here the mechanism has yet to be explained. In this series of papers, I present a detailed study of the origin of spiral activity in simulations in the hope that the mechanism that causes the patterns is also responsible for some of these features galaxies. In this first paper, I use a suite of highly idealized simulations of a linearly stable disk that employ increasing numbers of particles. While the amplitudes of initial non-axisymmetric features scale as the inverse square root of the number of particles employed, the final amplitude of the patterns is independent of the particle number. I find that the amplitudes of non-axisymmetric disturbances grow in two distinct phases: slow growth occurs when the relative overdensity is below ∼2%, but above this level the amplitude rises more rapidly. I show that all features, even of very low amplitude, scatter particles at the inner Lindblad resonance, changing the distribution of particles in the disk in such a way as to foster continued growth. Stronger scattering by larger amplitude waves provokes a vigorous instability that is a true linear mode of the modified disk.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Despite considerable effort over many decades, no theory for the origin of spiral patterns in galaxies has yet gained wide acceptance. There is good evidence (e.g., Kormendy & Norman 1979; Kendall et al. 2011) that some patterns in galaxies are excited by tidal interactions (e.g., Byrd & Howard 1992; Salo & Laurikainen 1993; Dobbs et al. 2010), others may be driven by bars (e.g., Salo et al. 2010), and perhaps even dark matter halo substructure (e.g., Dubinski et al. 2008). But when all these agents are excluded from N-body simulations, they still manifest spiral patterns and it therefore seems reasonable to expect that self-excited features also develop in galaxies.

Several reviews (Toomre 1977; Athanassoula 1984; Bertin & Lin 1996; Binney & Tremaine 2008; Sellwood 2010a) give detailed summaries of theoretical efforts to account for self-excited spirals. Most theorists agree that spirals are gravitationally driven density waves in the old stellar disk, a point of view that is strongly supported by the light distribution in the near-IR (e.g., Grosbøl et al. 2004; Zibetti et al. 2009; Kendall et al. 2011) and by streaming motions in high spatial resolution velocity maps (e.g., Visser 1978; Chemin et al. 2006; Shetty et al. 2007). However, strong disagreements arise over other aspects, in particular, the lifetimes of the spirals.

Small-amplitude, non-axisymmetric Jeans instabilities of equilibrium models are of two fundamental types. Cavity modes are standing waves, reminiscent of those in organ pipes and guitar strings, in which traveling waves reflect off an inner boundary, often the galaxy center, and corotation, where they are amplified. The bar-forming instability is the strongest mode of many simple models (e.g., Kalnajs 1978; Jalali 2007) and is of the cavity type (Toomre 1981). The other type of mode, such as the edge mode (Toomre 1981; Papaloizou & Lin 1989) or the groove mode (Lovelace & Hohlfeld 1978; Sellwood & Kahn 1991), are driven from corotation. All of these are vigorous instabilities that exponentiate on an orbital time scale.

Bertin & Lin (1996) argue that spirals are caused by a more gentle WASER instability (Mark 1977), a cavity mode that reflects off a "Q-barrier" instead of the center, that may become quasi-steady at finite amplitude through nonlinear effects. However, their specific examples (Bertin et al. 1989) of low-mass, cool disks that are needed to support slowly growing bisymmetric modes have been shown (Sellwood 2011) to develop more vigorous multi-arm instabilities that quickly alter the basic state they invoke.

Goldreich & Lynden-Bell (1965) and Toomre (1990) argue that, in galaxies lacking a bar or a recent interaction with a companion, spirals are no more than shearing bits and pieces of swing-amplified noise. Substantial inhomogeneities, such as star clusters and giant molecular clouds, raise the noise level far above that expected from the stars alone, and the noise level can be amplified ∼100 fold in cool, responsive disks. However, since the spirals they wish to account for themselves cause the disk to heat quickly, it is hard to understand how a galaxy could sustain the combination of responsiveness and noise level needed to create spiral features of the amplitude we observe. Furthermore, this mechanism has difficulty accounting for even a modest degree of symmetry in the patterns (e.g., Kendall et al. 2011).

From the earliest days (Miller et al. 1971; Hockney & Brownrigg 1974; James & Sellwood 1978), N-body simulations of dynamically cool collisionless disks have manifested, recurrent short-lived transient spirals. Claims of long-lived waves in some simulations are not reproducible (Sellwood 2011). Sellwood & Carlberg (1984) showed that spiral activity in purely stellar disks is self-limiting because potential fluctuations caused by the rapidly evolving spirals themselves scatter stars away from circular orbits; the increased random motion in the disk reduces its ability to support further collective oscillations. Sellwood & Carlberg (1984) also found that spiral activity can continue "indefinitely" in models that mimic the effects of gas dissipation and star formation, offering an attractive explanation for the observed coincidence between spiral activity and gas content in galaxies. The behavior they reported continues to occur in modern simulations of galaxy formation (e.g., Abadi et al. 2003; Roškar et al. 2008; Agertz et al. 2011).

Local theory (Julian & Toomre 1966) predicts that a perturbing mass moving on a circular orbit in a self-gravitating disk surrounds itself with a wake. In an N-body simulation, each particle is itself a perturber and has its own peculiar motion that is typically as large as that of the surrounding particles whose motion it affects. Note that a system of dressed particles, with each orbiting at its own angular rate, will give rise to shearing density fluctuations that will be strongest when the wakes are aligned, giving the appearance of swing amplification. In fact, the picture of randomly placed dressed particles in a shearing disk is equivalent to that of swing-amplified noise, with the input noise level raised by the mass cloud surrounding each source particle (Toomre 1990).

Toomre & Kalnajs (1991) studied swing-amplified particle noise using simulations in shearing (x, y)-coordinates in a periodic box. They had to counter the continuous rise in the level of random motion in order to be able to study the long-term equilibrium level of spiral activity, and they therefore added a "dashpot" type damping term to the radial acceleration, fx = −Cvx. The level of random motion should have settled to a roughly constant value if the damping constant C were proportional to the inverse of the particle number. They found, however, that the observed rms velocity of the particles exceeded their predicted level because the density fluctuations were about twice as large as they expected—a discrepancy that they attributed to additional correlations between the particles that developed over a long period.

While the origin of spiral features in global simulations has yet to be explained, their amplitude in larger N-body simulations seems to be too great to be just swing-amplified particle noise (Sellwood & Carlberg 1984). Fujii et al. (2011) found that spirals in simulations with larger numbers of particles heated the disk more slowly, allowing the spiral activity to continue for longer. These authors incorrectly attributed the rapid heating rate in the smaller N simulations by Sellwood & Carlberg (1984) to two-body relaxation, whereas the higher level of shot noise simply provokes stronger spirals.1

In this paper, I attempt to uncover the reason for the suprisingly vigorous spiral activity in N-body simulations of cool, collisionless disks. The models employed here are highly idealized in order to simplify the dynamics to the point that it becomes understandable. Improved realism is deferred to later papers. I have long hoped (Sellwood 1991, 2000) that, in realistic galaxy models, the decay of one spiral feature might change the background disk in such a manner as to create conditions for a new instability, much as Sellwood & Lin (1989) had observed in simulations of a cool, low-mass mass disk having a near-Keplerian rotation curve. Here I finally present the evidence that something similar to this idea really does occur in more galaxy-like models, but the behavior differs in detail from that which I envisioned.

2. A LINEARLY STABLE DISK

The disk in which the circular speed is independent of radius, now known as the "Mestel disk" (Binney & Tremaine 2008), is predicted by linear stability analyses (Zang 1976; Toomre 1981; Evans & Read 1998) to be globally stable to most non-axisymmetric disturbances. I therefore use this model in the simulations described in this paper.

The potential in the disk plane is

Equation (1)

where V0 is the circular speed and Ri is some reference radius. The surface density of the full-mass disk is

Equation (2)

A star at radius R moving in the plane with velocity components vR and vϕ in the radial and azimuthal directions, respectively, has specific energy E = Φ(R) + (v2R + vϕ2)/2 and specific angular momentum Lz = Rvϕ. A distribution function (DF) for this disk that yields a Gaussian distribution of radial velocities of width σR is (Toomre 1977; Binney & Tremaine 2008)

Equation (3)

where q = V20R2 − 1 and the normalization constant

Equation (4)

In order to avoid including stars having very high orbital frequencies near the center, Zang (1976) introduced a central cutout in the active surface density by multiplying the distribution function f by the taper function

Equation (5)

where the index ν determines the sharpness of the taper. This taper function removes low angular momentum stars, reducing the active disk mass at the center, which must be replaced by unresponsive matter in order that the total potential is unchanged. The characteristic radius Ri of this cutout introduces a scale length. Henceforth, I adopt units such that V0 = Ri = G = 1.

Zang (1976) found that this almost full-mass disk was stable to all bisymmetric modes, provided the index ν ⩽ 2. This important result was confirmed by Evans & Read (1998), who also found that other disks with power-law rotation curves were remarkably stable to bisymmetric modes. We now understand that this disk model is linearly stable to bar-forming modes through Toomre's (1981) mechanism because the frequency singularity at the disk center ensures that disturbances of any pattern speed Ωp will have an inner Lindblad resonance (hereafter ILR) as long as m > 1.

However, Zang also found that the full-mass Mestel disk was globally unstable to lopsided (m = 1) modes, as are the power-law disks (Evans & Read 1998). This instability persists no matter how hot the disk or gentle the central cutout. The mechanism for the lopsided mode is also a cavity mode, since feedback through the center is allowed only for m = 1 waves because the ILR condition in power-law disks cannot be satisfied when Ωp > 0.

Swing amplification in a disk having a flat rotation curve is most vigorous for 1 ≲ X ≲ 2.5 (Julian & Toomre 1966; Toomre 1981), where X = Rkcrit/m, with the smallest wavenumber for axisymmetric instabilities kcrit = κ2/(2πGΣ) (Toomre 1964; Binney & Tremaine 2008). Since X = 2/m in a full-mass Mestel disk, the swing amplifier is at its most vigorous for m = 1 disturbances. The parameter X = 2/(ξm) is increased by reducing the active surface density of the disk to ξΣ without changing the circular speed; here 0 ⩽ ξ ⩽ 1. Since the swing amplifier is all but dead when X > 3 in a flat rotation curve, the lopsided instability should disappear for ξ ≲ 2/3.

Toomre (1981) reports that a half-mass (ξ = 0.5) disk, with ν = 4 for the central cutout, has no small-amplitude, global instabilities for any m. This model is the only known globally stable disk with differential rotation and significant disk mass.2

As the disk still extends indefinitely to large R, A. Toomre (1989, private communication) added an outer taper

Equation (6)

with index μ to control the abruptness of the taper about the angular momentum characteristic of a circular orbit at some large radius Ro. Again, the potential is unchanged if the mass removed is replaced by rigid matter. He found μ ⩽ 6 to be adequate to avoid provoking outer edge modes (Toomre 1981; Papaloizou & Lin 1989).

As this taper function still does not limit the radial extent of the disk, but merely reduces the active density of the outer disk, I have to add the additional constraint that f(E, Lz) = 0 when E > Φ(Rmax), so that the disk contains no particles with orbits that extend beyond Rmax. In order that this additional limit does not also provoke an edge mode, I choose Rmax ≳ 1.7Ro, so that it removes mass only where the disk density is already very low.

In the simulations reported in this paper, I adopt ξ = 0.5 and q = 11.44 so that the half-mass disk has nominal Q = 1.5, the inner cutout is characterized by ν = 4 at Ri = 1, and the outer taper has μ = 5 at Ro = 11.5 and the limiting radius is Rmax = 20.

Figure 1 shows the density of particles in the space of the two actions JR and Jϕ (Binney & Tremaine 2008). In an axisymmetric two-dimensional disk model, the azimuthal action JϕLz, while the radial action,

Equation (7)

expresses the amount of radial motion the particle possesses. Here, vR = [2E − 2Φ (R) − L2z/R2]1/2 (positive root only in this definition) and the limits of integration are the radii of peri- and apocenter of the orbit where the argument of the square root is equal to zero. Thus, JR also has the dimensions of angular momentum and JR = 0 for a circular orbit. The actions are an alternative set of integrals to the classical integrals E and Lz, but the relation between the two sets of integrals is cumbersome.

Figure 1.

Figure 1. Initial density of particles in the space of the two actions. The contour spacings are linear from 5% to 95% of the function maximum. The density for low values of Jϕ is suppressed by the inner angular momentum taper and, as is generally true, the density is greatest for near circular orbits (JR = 0) and decreases for more eccentric orbits.

Standard image High-resolution image

3. SIMULATIONS OF THE "STABLE" DISK

The model just described, with the outer taper but not the energy cutoff, was predicted by A. Toomre (1989, private communication) to have no small-amplitude instabilities. Here I report simulations with this model both to test the prediction and to study how replacing the infinitely finely divided "stellar fluid" by a finite number of particles affects the behavior.

I use the two-dimensional polar-grid code (Sellwood 1981) with multiple time step zones (Sellwood 1985), which has previously been shown to reproduce the predicted normal modes of unstable disks (Sellwood & Athanassoula 1986; Earn & Sellwood 1996; Sellwood & Evans 2001). Gravitational forces between particles obey a Plummer softening law, with the softening parameter epsilon = Ri/8, so that forces can be thought of arising from a disk of thickness epsilon.

The grid has 128 spokes and 106 radial rings, and I restrict disturbance forces to the m = 2 sectoral harmonic only; the axisymmetric radial force is −V20/R at all radii. I employ a basic time step of δt = Ri/(40V0), which is increased by successive factors of two for particles with radii R > Ri, R > 2Ri, R > 4Ri, and R > 8Ri. Time steps are also sub-divided when particles move near the center in a series of "guard zones" (Shen & Sellwood 2004), without redetermining the gravitational field, which is an adequate approximation since the dominant force in this region arises from the rigid component. Direct tests showed that the results are insensitive to moderate changes of grid size or time step.

This paper reports results from several simulations that differ in the number of particles, and I label each run by the number of particles. Much of the analysis focuses on run 50M, which therefore has 5 × 107 particles, and its variants runs 50Ma, 50Mb, etc.

4. RESULTS

Figure 2 shows the evolution of the relative overdensity of bisymmetric disturbances in a series of simulations with increasing numbers of particles. The upper panel shows the quantity δmax, which is the largest value of the ratio δΣ/Σ over the radial range 1.2 < R < 12, i.e., the part of the disk where the surface density has not been significantly reduced by the tapers. The unit of time is Ri/V0 and, since I have chosen units such that Ri = V0 = 1, the orbit period at radius R is 2πR.

Figure 2.

Figure 2. Upper panel: the time evolution of the peak overdensity in a series of simulations of the half-mass Mestel disk with different numbers of particles. The model is predicted by A. Toomre (1989, private communication) to be globally stable. The ordinate reports the maximum value of δΣ/Σ on grid rings over the range 1.2 < R < 12 where the surface density is little affected by the tapers. Linear theory predicts the amplitude should remain proportional to N−1/2. Lower panel shows the radius of the measurement reported in the upper panel for run 500M only. The apparent lower bound is because values for R < 1.2 were excluded to eliminate shot noise in the tapered part of the disk.

Standard image High-resolution image

As more particles are employed, the initial value of δmax decreases as N−1/2, as it must for randomly placed particles. Its value rises by a factor of a few in the first few dynamical times as the initial shot noise is swing amplified. In the smallest N experiment, the initially amplified noise is already close to the saturation level, but as N rises, the amplitude takes increasingly long to reach its maximum. However, the final amplitude is independent of N for all particle numbers up to N = 5 × 108. In the larger N experiments, the amplitude increase is characterized by roughly exponential growth at two distinct rates: an initial period of slow growth, followed by a steeper rise once δmax ≳ 0.02. These two separate rates of growth are approximately independent of N over the amplitude ranges where they are observed.

It might seem that exponential growth indicates an unstable normal mode, or perhaps a few such instabilities, but such an interpretation is unattractive for a number of reasons. Simulations of linearly unstable models (e.g., Sellwood & Athanassoula 1986; Earn & Sellwood 1996) usually reveal the mostly vigorous instability emerging from the noise at an early stage and dominating the subsequent growth until it saturates. Modulated growth could occur in a system that supports a small number of overstabilities having similar growth rates but different pattern speeds, since the changing relative phases of the modes over time causes beat-like behavior. In all such cases, however, the overstabilities should maintain phase coherence until they saturate, but power spectra of these simulations (shown below) reveal multiple features none of which retains phase coherence throughout the period of growth. Furthermore, were the later, more rapid rise due to the emergence of more rapidly growing normal modes of the original disk, it is unusual that they should take so long to rise above the amplitude of the more slowly growing "modes" that would also need to be invoked to account for the initial slow rise, and it is most unlikely that the change of slope would occur at almost the same amplitude, as suggested for the three larger N experiments (Figure 2). Finally, Toomre's (1981) linear stability analysis that predicted the model to be globally stable also argues against this interpretation of the results.

4.1. Slow Growth

The lower panel of Figure 2 shows that the maximum disturbance density in run 500M is generally in the inner disk for t < 2500, i.e., until the saturation amplitude is reached. The behavior over the entire early period 0 < t < 2000 seems to follow a quasi-repetitive pattern: a density maximum appears in the range 3 ≲ R ≲ 6 that then propagates inward. This pattern results from swing-amplified noise creating a trailing spiral disturbance that travels inward at the group velocity, as shown in the "dust to ashes" figure of Toomre (1981). Salo & Laurikainen (2000, their Figure 7) report very similar behavior in their three-dimensional simulations of a low-mass exponential disk embedded in a rigid halo.

The gradual rise of spiral activity in the global simulations reported here can also be viewed as the buildup of mass clouds surrounding each particle. Figure 3 shows that each particle can be regarded as being "dressed" by a spiral wake, as in the local context. To construct this figure at each instant, I stacked copies of the grid-estimated density distribution from run 50M by scaling the entire grid radially and rotating it such that the density maximum on each "source" ring in turn lay at R = 7 and zero azimuth. At the initial moment, the non-axisymmetric part of the combined density shows no features other than the azimuthal extension of the source caused by forcing cos 2θ angular dependence. As time progresses, each source particle becomes dressed by a wake of gradually increasing mass and spatial scale; the maximum overdensity shows a general rise, with fluctuations that closely follow the time variations of δmax in the same simulation, shown by the cyan line in Figure 2.

Figure 3.

Figure 3. Gradual development of polarization in the mass distribution of run 50M. The contours show the positive part only of the disturbance density, with the bisymmetry enforced by the code, and contour levels are the same in each panel. The inner boundary is at R = 1.5, the outer at R = 15, and the "source" particles are at R = 7. The peak relative overdensity at each time is reported in the lower right corner.

Standard image High-resolution image

However, two related aspects of the behavior are especially noteworthy. First, growth in the large N simulations continues for a much longer period than expected. Julian & Toomre 1966 found, from a local linear analysis, that the wake takes ∼5 epicycle periods to become fully developed. Swing-amplification of particle shot noise causes the immediate rise in the first few time units, as already noted, which is a large part of the development of spiral wakes. The epicyclic period at radius R in the Mestel disk is 21/2πR/V0, or ∼44 time units at R = 10, about halfway out in the disk. Thus the period of continued gradual growth greatly exceeds five epicycle periods at a typical radius, which is the timescale expected from linear theory. Second, the simulations do not reveal a limiting amplitude; instead more rapid growth takes over in every case, even when 5 × 108 particles are employed. Section 4.6 presents further analysis of the slow-rise phase that accounts for these differences.

4.2. Rapid Rise

The acceleration in the rate of growth once δmax ≳ 0.02 is clearly inconsistent with the linear theory prediction. The enhanced growth rate is again approximately independent of N, as is the final amplitude at a relative overdensity between 20% and 30%.

Figure 4 shows a number of power spectra, i.e., contours of power as functions of radius and frequency, from run 50M. A horizontal ridge in these figures indicates a coherent density disturbance that extends over a range of radii and has an angular frequency mΩp, with m = 2 and Ωp being the rotation rate or pattern speed. There is very little power at frequencies higher than those shown. The solid line indicates the frequency of circular motion, mΩ and the dashed lines mΩ ± κ. Each panel gives the spectrum over an interval of 200 time units, with the start times of successive panels shifted forward by 100 time units.

Figure 4.

Figure 4. Power spectra of m = 2 disturbances in run 50M. Successive contour levels, which start from the same value in each plot, differ by factors of two. Each panel is taken from data over the period indicated, which overlap so that only half the plots are from independent data. The solid curves show 2Ω(R) and the dashed curves 2Ω(R) ± κ(R). Data outside the radial ranges marked by vertical dotted lines, where the surface density is low, are excluded because they are too noisy.

Standard image High-resolution image

As shown in the lower panel of Figure 2 for run 500M, density fluctuations at early stages in run 50M also arise at radii 4 ≲ R ≲ 7 from swing-amplified shot noise and subsequently propagate inward at the group velocity (Toomre 1969). The amplitude of each disturbance rises as it approaches the ILR because wave action is conserved as the group velocity decreases until wave–particle interactions absorb it near the ILR (Lynden-Bell & Kalnajs 1972; Mark 1974).

The behavior changes around t ∼ 1300. The strongest feature up to this time has a frequency ∼0.5 (e.g., third panel of Figure 4), but thereafter a new much stronger disturbance appears with a frequency ∼0.55. Analysis with the mode-fitting apparatus described in Sellwood & Athanassoula (1986) using data from the period 1100 ⩽ t ⩽ 1500 reveals that this wave has a frequency ω = mΩp + iγ ≃ 0.55 + 0.13i, although the coexistence of other waves over the same period makes this frequency estimate somewhat uncertain. The significant growth rate suggests it is a true instability, yet it is scarcely conceivable that this mode could have been growing at this rate from t = 0, since that would require its initial amplitude to have been smaller by ≳ 50 e-folds before it first became detectable.

4.3. Randomized Restarts

Figure 5 reports the results of four further simulations that were restarted from rearrangements of the particle distribution at selected times from run 50M. The original run is reproduced as the cyan curve.

Figure 5.

Figure 5. Time evolution of the peak overdensity in five simulations with N = 50M. The cyan curve is for run 50M, reproduced from Figure 2, while the green, red, and magenta curves show the behavior of runs 50Ma, 50Mb, and 50Mc created by randomizing the azimuths of the particles at t = 1000, 1200, and 1400, respectively. For the blue curve, run 50Md, the particles preserved the same E and Lz, but the radial phase was also reset. Both runs restarted from t = 1400 manifest a single dominant instability.

Standard image High-resolution image

To construct three of these simulations, I simply added to the azimuthal phase a random angle selected uniformly from 0 to 2π, while leaving the radius and velocity components in cylindrical polar coordinates unchanged. This procedure erases any non-axisymmetric feature above the shot noise level, and the starting amplitude is similar to that of the cyan line at t = 0.

The green, red, and magenta curves, respectively, show the evolution of runs 50Ma, 50Mb, and 50Mc, which were re-started from t = 1000, t = 1200, and t = 1400. In all cases the peak non-axisymmetric density rises more quickly than that of the original run over the period 200 ≲ t ≲ 1000. Furthermore, the later the restart time, the more rapidly the amplitude rises and the sooner the amplitude saturates! The different behavior of these randomized runs can only be a consequence of the changes to the particle distribution that occurred in the original run prior to azimuthal randomization.

The most dramatically different behavior is that of run 50Mc (magenta), which manifests a single vigorous instability of eigenfrequency ω = 0.6 + 0.16i—close to that of the dominant disturbance around t = 1400 identified in the earlier analysis of run 50M (Section 4.2). The form of the unstable mode is illustrated in Figure 6. Clearly, the particle distribution at t = 1400 has a feature that provokes this instability, and which must have been created by the earlier evolution.

Figure 6.

Figure 6. Best-fit mode extracted from the period of exponential growth in run 50Mc using the procedure described by Sellwood & Athanassoula (1986). The circles mark the radii of the principal resonances.

Standard image High-resolution image

The behavior of the other two cases (green and red curves of Figure 5) differs not quite so markedly from the original, but again indicates that the particle distribution at both t = 1000 and t = 1200 contains features that lead to an accelerated rise of non-axisymmetric features. I return to these cases in Section 4.6.

Note that randomizing the azimuthal phases at t = 1400 only made the identifiable instability of the original model stand out more clearly. Thus, the destabilizing agent cannot be a previously created non-axisymmetric feature, e.g., a weak bar, in the particle distribution, and must therefore be a feature in the distribution either of the radial phases, or of the integrals (E and Lz or JR and Jϕ).

In order to test whether changes to the distribution of radial phases are important, I determined E and Lz for each particle and then selected at random a new pair of radial and azimuthal phases from an orbit having these integrals in the analytic potential of the Mestel disk. These new phases determine new positions and velocities for each particle, although it has the same integrals as before. The evolution in this case, run 50Md shown by the blue line in Figure 5, makes it clear that the same instability is present as that in run 50Mc (magenta line). Thus, the important change by t = 1400 of the original run is to the distribution of the integrals.

4.4. Changes to the Distribution Function

Figure 7 shows the distribution of particles in action space of run 50M at the times t = 1200 (upper) and t = 1400 (lower). Although the contours are noisier, caused by a degradation of the original smooth arrangement, their overall shape in the upper panel is little changed from that at t = 0 (Figure 1). However, more substantial changes are visible in the lower panel. The changes are shown more clearly in the upper panel of Figure 8, which contours differences between the distributions at t = 1400 and at t = 0, with increases shown by the blue contours and decreases by the red contours.

Figure 7.

Figure 7. Action-space density of particles in run 50M at t = 1200 (upper) and t = 1400 (lower).

Standard image High-resolution image
Figure 8.

Figure 8. Upper panel contours the differences between the action-space density of particles in run 50M at t = 1400 and t = 0, with positive (negative) differences being indicated by blue (red) contours. The dashed lines show the loci of Lindblad resonances (cyan) and of corotation (green) for Ωp = 0.25, while the solid lines indicate the scattering trajectory from JR = 0 for the Lindblad resonances. The lower panel shows the mean displacements of particles in the same space and over the same period. The arrow size is proportional to the mean value of the displacement in each element.

Standard image High-resolution image

The various lines in Figure 8 show resonance loci and scattering trajectories for a disturbance of angular frequency mΩp = 0.5, which is the frequency of the dominant feature of the power spectrum shown in the third panel of Figure 4. The resonance lines (dashed) are computed for Ωp = Ωϕ + lΩR/m, where l = 0 for corotation and l = ∓1 for the inner and outer Lindblad resonances (OLRs), and the frequencies Ωϕ and ΩR are computed for orbits of arbitrary eccentricity. The scattering trajectories (solid lines) are computed assuming ΔE = ΩpΔLz, as required by the conservation of the Jacobi constant in a rotating non-axisymmetric potential (Binney & Tremaine 2008), with ΔE converted to ΔJR and assuming JR = 0 initially for a particle in resonance. Note that any scattering at corotation would occur with ΔJR = 0.

It is clear that the principal changes to the distribution of particles in action space were caused by scattering at the ILR of the strongest non-axisymmetric feature to have developed before t = 1400. This conclusion is considerably strengthened by the evidence in the lower panel of Figure 8, which shows the mean displacements of particles in action space between t = 0 and t = 1400. This figure not only confirms the convergence of vectors toward the density maximum created by scattering at the ILR, but also reveals that substantial changes have taken place both at corotation and at the OLR, which produce no net change in density. Note that the vectors along the Lindblad resonance lines are parallel to the scattering directions for those resonances, computed for the same pattern speed, while they are near horizontal at corotation. Thus, particles at all three resonances are scattered with no change to Jacobi's invariant for the measured pattern speed. The displacements at the OLR cause no apparent net change to the density (upper panel), while those at the ILR do. The different behavior at the two Lindblad resonances is partly caused by the geometric consequence of the "focusing" of waves that propagate inward, while outgoing waves spread out over an increasing disk area. It is also significant that the scattering direction at the ILR is very close to the resonance locus, so that particles stay on resonance as they lose angular momentum, whereas a small gain or loss of angular momentum at the OLR moves the particle off resonance.

There are a few additional features visible in the lower panel of Figure 8 that appear to be due to another somewhat lower frequency wave that is also present at t = 1400 (see the fourth panel of Figure 4). These features are much weaker in an equivalent plot (not included here) of the lower panel of Figure 8 constructed for t = 1300.

Scattering at the ILR naturally causes a very localized increase in the random motions of the particles. Figure 9 shows the radial variation of the second moment of the radial velocities estimated from the particles. By construction from the analytic DF  the radial velocity distribution is near Gaussian at t = 0, but the resonant scattering in action space does not preserve this property. Note that this very localized heating at a resonance was caused by a disturbance that had a relative overdensity of ∼2%.

Figure 9.

Figure 9. Second moment of the radial velocity distribution of particles in run 50M at the three indicated moments. No visible changes have occurred by t = 700, while a significant increase can be seen near the center at t = 1400.

Standard image High-resolution image

4.5. Nature of the New Instability

What causes the new instability that appears in runs 50Mc and 50Md, and which is also present, but less distinct, in the continued run 50M? Since randomizing the azimuthal coordinates makes the instability stand out more clearly, it cannot be caused by trapping of particles into some non-axisymmetric feature; neither can it be caused by the kind of nonlinear wave coupling discussed by Masset & Tagger (1997) or Fuchs et al. (2005). Its rapid exponential growth characterizes it as a vigorous, global, linear mode of the modified disk.

It has long been my expectation (e.g., Sellwood 1991, 2000) that the recurrent cycle would prove to be one of groove-type modes (Sellwood & Kahn 1991) caused by depopulating the DF near the OLR of one wave to provoke a new instability with corotation near the previous OLR. Indeed this is what Sellwood & Lin (1989) observed in their simulations of a low-mass disk with a near Keplerian rotation curve. Yet I have been unable to find evidence to support this expectation in more massive disks with flatter rotation curves, and the new evidence presented in Figure 8 seems quite emphatically to rule it out, at least for this case. While changes to the DF are clearly responsible, it is not scattering at the OLR, and some other mechanism is at work.

The only significant change to have occurred to cause runs 50Mc and 50Md to behave so differently from the initial behavior of run 50M is the localized heating near R = 1.17, which is the radius of the ILR for near circular orbits for the scattering wave. There are no net changes elsewhere, even though the lower panel of Figure 8 indicates that some particles in other parts of the disk have changed places.

Since the new instability has a slightly higher pattern speed than that of the preceding scattering wave, the radii of the principal resonances are all correspondingly smaller, and therefore the ILR of the new instability would lie within the heated region of the disk. The localized deficiency of nearly circular orbits created by the previous changes to the DF will have strongly inhibited a supporting response in this region of the disk to forcing by an inward propagating density wave.

Thus, it seems to me that the most plausible unstable mode created by these "initial" conditions is a cavity-type mode that is a standing wave between corotation and a hard inner reflection off the region of the disk where the DF has been changed. Since the inner disk is "hotter," there is a superficial resemblance between this suggestion and the WASER mechanism proposed by Mark (1977) and advocated by Bertin & Lin (1996). However, the WASER mode has a low growth rate because the out- and ingoing waves at corotation are both trailing, resulting in very mild amplification as the disturbance connects only the long- and short-wave branches of the dispersion relation. Furthermore, their mode requires a low surface mass density and Q ∼ 1.0 near corotation, whereas this model has a massive disk with Q = 1.5. Finally, the change in disk properties that causes the inner reflection is quite different from the simple increase in Q postulated for the WASER mechanism. Much more vigorous modes, such as that observed here, seem likely to require a full swing from leading to trailing (Toomre 1981). Thus, I suspect that the ingoing trailing wave reflects off the region where the DF has been changed as an outgoing leading wave. I have no evidence that a hard reflection can occur in this case, so this proposed mechanism is entire speculation.

4.6. Earlier Restarts

Simulations 50Ma and 50Mb were restarted at t = 1000 and t = 1200 respectively. As shown in Figure 5, density fluctuations in these runs grew at rates that were intermediate between that of the original run 50M and the clear instability exhibited by those (50Mc and 50Md) that were restarted at t = 1 400. Moreover, these two runs support multiple disturbances having several different angular frequencies, none of which maintains coherence through to the saturation amplitude. The randomized particle distributions at these intermediate times must also differ in important ways from the smooth particle distribution of run 50M at t = 0.

Figure 10 shows that the changes to the DF by t = 1000 are tiny, but nevertheless significant. I had to reduce the contour levels in order to reveal the tiny changes to the density of particles in action space (top panel), which necessitated a proportionate increase in the kernel size so that the shot noise from the number of particles contributing to the lowest contour is unchanged. It is clear that there has been some significant scattering of particles away from the JR = 0 axis, especially at small radii. These, and the significantly larger changes that occur by t = 1200 (not shown) caused some heating of the innermost part of the disk, as illustrated in the lower panel. This very mild heating of the inner disk is caused by the absorption at the ILRs of incoming disturbances created by swing-amplified shot noise at larger radii. Although the wave amplitudes were tiny, they caused a lasting and significant change to the DF.

Figure 10.

Figure 10. Upper panel contours the differences between the action-space density of particles in run 50M at t = 1000 and t = 0, with positive (negative) differences being indicated by blue (red) contours. The kernel width is five times larger than that used in Figure 8 and the contour levels are one-fifth. The lower panel shows the radial velocity dispersion of particles in run 50M at the three indicated moments. The changes from t = 0 are non-zero only near the center; they are tiny by t = 1000 and moderately larger at t = 1200.

Standard image High-resolution image

The randomized restarts at t = 1000 (run 50Ma) and t = 1200 (run 50Mb) confirm that these tiny changes, and not nonlinear coupling, say, were responsible for the enhanced growth of non-axisymmetric structure. It seems likely that the small changes that have taken place are sufficient to cause some partial reflection of waves incident on the inner disk that boosts the density fluctuations, but is apparently insufficient to provoke an indefinitely growing, more vigorous mode at these times.

Thus, the slow rise in the density fluctuations described in Section 4.1 occurred because the physical system gradually develops a marginally modified inner disk. Mild scattering of particles in the inner disk as the wave action of the noise-driven structures created at larger radii is absorbed at the appropriate ILR. These changes to the dynamical properties of the inner disk seem to be responsible for gradually increasing partial reflection of later incoming waves, and this partial feedback in turn enhances the amplitudes of subsequent disturbances. Growth is slow until the inner disk reflects waves strongly enough to create a coherent linear instability.

Note that the slow, fluctuating rise in δmax in both runs 50M and 500M (upper panel of Figure 2) is roughly exponential, with similar exponents in both cases. This appears to suggest that the behavior just described is, in fact, destabilizing in the sense that it implies that the rate of growth is independent of the amplitude. It would seem, therefore, that no system of particles, however large, could behave as a smooth disk.

5. SUMMARY AND DISCUSSION

The main result presented here is the demonstration that scattering at the ILR of a spiral disturbance of even very low amplitude causes a lasting change to the properties of the disk that leads to increased amplitude of subsequent activity. Linear perturbation theory could never capture this behavior, because it neglects second-order changes to the equilibrium model caused by a small amplitude disturbance.

The random density fluctuations of a system of self-gravitating particles in a cool, shearing disk are amplified as they swing from leading to trailing (Toomre 1990; Toomre & Kalnajs 1991). The wave action created by these collective responses is carried inward at the group velocity to the ILR where it is absorbed by particles. The resulting scattering of particles to more eccentric orbits de-populates the near-circular orbits over a narrow range of initial angular momentum. Here I have shown that, no matter how small the amplitude, the lasting changes to the particle distribution promote a higher level of density fluctuations, which leads to indefinite growth.

The growth of non-axisymmetric waves in the idealized simulations presented here is characterized by two phases. When very large numbers of particles are employed, the behavior appears to be that of a system of dressed particles (Toomre & Kalnajs 1991; Weinberg 1998), but one in which the amplitudes of the particle wakes rise slowly. As the peak relative overdensity passes ∼2%, the changes to the background state are destabilizing and simple instabilities appear that exhibit runaway growth.

The instability that appears is caused by earlier changes to the DF and is a true mode that could have grown exponentially from low amplitude. Direct tests reported in Section 4.3 demonstrate that it does not rely upon nonlinear coupling to previous structures. I speculate on a possible mechanism for the unstable mode in Section 4.5.

Linear perturbation theory (Toomre 1981) predicts no instability for a smooth DF, yet the evidence from Section 4.6 is that no finite number of particles would ever avoid indefinite growth. Whether or not this conclusion is correct, the avoidance of runaway growth for some non-infinite number of particles is of theoretical interest only, as galaxies are probably never as smooth as a disk of 5 × 108 randomly distributed particles that are each only ∼10 times as massive as a typical star.

Similar evolutionary changes to the original smooth disk are undoubtedly the cause of the non-axisymmetric features that developed for m > 2 in the simulations reported by Sellwood (2011) to test the models proposed by Bertin et al. (1989). The appearance of visible multi-arm waves that heated the disk in those models was, as here, more and more delayed as the number of particles was increased. Note that no large amplitude waves developed when forces were restricted to m = 2 because the disk was of such low mass. The swing amplification of shot noise is minimal when the parameter X > 3, which is the case for bisymmetric waves in this low-mass disk, and resonant scattering by such weak disturbances will cause extremely slow growth, if any. However, Xm−1, making the disk more responsive to disturbances with m > 2, allowing runaway growth of multi-arm features and disk heating once these force terms were included.

Toomre & Kalnajs (1991) used local simulations in a small shearing patch of a disk to study the behavior of swing-amplified shot noise. They were able to explain even the long-term behavior with just linear theory, although the limiting amplitude was a factor of ∼2 higher than they estimated, which they attributed to additional correlations between the positions of particles that developed in the long term. Most significantly, they did not observe the kind of runaway growth reported here. This difference may be due to a combination of factors: first, the particles in their principal box interacted with pairwise forces and, had the box extended infinitely in the azimuthal direction, each non-axisymmetric feature would be composed of a continuous spectrum of waves, implying that there would be no equivalent of Lindblad resonances and the absorption of wave action would be spread over a broad swath of particles. However, their periodic boundary conditions implied that the spectrum of waves was discrete, and a Lindblad resonance must have been present for each of the multiple waves that made up an individual shearing feature. Thus, the absorption of wave action must have taken place at a number of discrete locations, and mild resonant scattering probably occurred at each. But the most important distinguishing feature of their simulations that prevented development of destabilizing features at these resonances is that they applied a damping term to only the radial velocity component of their particles. Any resonant scattering that depopulated nearly circular orbits must have been quickly erased, preventing their system from manifesting the behavior described here.

I am grateful to the referee for suggesting that I note the observations of Saturn's rings (reviewed in Cuzzi et al. 2010), which seem to support the picture of spiral chaos proposed by Toomre (1990). Of course, the snowball particles in this beautiful ring system are believed to collide dissipatively about once per orbit, which would again prevent scattering at possible Lindblad resonances from creating reflecting regions that might destabilize the disk.

While identifying the origin of indefinite growth is a good beginning to a detailed understanding the origin of spiral features in N-body simulations, it is far from being the whole story. All the simulations reported here supported several coexisting waves of differing frequencies at all times, except runs 50Mc and 50Md in which pre-existing waves were eliminated and single disturbance stood out for a brief period. The recurrence of large-amplitude waves remains to be described in a later paper of this series.

In order to be able to reach the level of understanding described here, I have simplified the dynamics of a realistic disk galaxy to the extreme. First, the particles in the simulations presented here interacted with each other only through the m = 2 sectoral harmonic of the disturbance forces, with all other self-gravity terms being ignored; even the mean central attraction was replaced by that of a rigid mass distribution. Second, the particles were constrained to move in a plane only. Third, I deliberately chose to study an idealized model that was known (Zang 1976; Toomre 1981; Evans & Read 1998) to possess no small-amplitude instabilities. Fourth, I selected particles so that their joint distribution of E and Lz matches that of the adopted DF as closely as possible. Fifth, the simulations were entirely isolated from external perturbations by passing satellites, halo substructure, etc., and the halo was represented by an idealized axisymmetric and time-independent mass distribution. Sixth, all effects of gas dynamics, such as the creation of dense clouds and the formation of stars, were excluded.

Yet even with all these abstractions, the behavior took some effort to understand. The demonstration of a simple, clean instability was possible only in restarted simulations, constructed by careful reshuffling of the particles at the crucial time. Simulations that include all the above-mentioned complications continue to manifest recurrent spiral patterns. It is reasonable to hope that their origin is related to the mechanism presented here, but this remains to be shown.

The ultimate goal of this work is to demonstrate that some spirals in galaxies are self-excited instabilities with a similar origin to that in the simulations. Sellwood (2010b) and Hahn et al. (2011) have made a step in this direction, by finding a feature in the distribution of solar neighborhood stars that resembles that of scattering at an ILR. Further analysis of data from other ground- and space-based surveys, such as HERMES (Bland-Hawthorn et al. 2010), APOGEE (Majewski et al. 2010), and especially Gaia (Perryman et al. 2001), will afford better tests.

I thank Alar Toomre for providing the parameters of the linearly stable model and for years of general encouragement to track down the source of the spirals in simulations. I also thank the referee for a careful report, Scott Tremaine for some suggestions to improve the presentation, and Tad Pryor for many helpful discussions. This work was supported by grants AST-0507323 and AST-1108977 from the NSF.

Footnotes

  • If collisional relaxation were rapid, it would have been impossible for the same code to have supported the global modes (Sellwood & Athanassoula 1986; Earn & Sellwood 1996; Sellwood & Evans 2001) predicted by linear perturbation theory for a collisionless stellar fluid, or to match the results from direct integration of the collisionless Boltzmann equation (Inagaki et al. 1986). The formula for the relaxation rate in a two-dimensional disk of particles does not contain the Coulomb logarithm (Rybicki 1972) and the rate is therefore suppressed more effectively by gravity softening.

  • Two of the composite Ω models reported by Kalnajs (1972) appear to be the only other known stable disks, but are of less interest because they rotate uniformly.

Please wait… references are loading.
10.1088/0004-637X/751/1/44