Next Contents Previous

2. TORUS CONSTRUCTION

Regular motion is defined as motion that respects at least N isolating integrals, where N is the number of degrees of freedom (DOF), i.e. the dimensionality of configuration space. Regular motion can always be reduced to translation on a torus; that is, a canonical transformation (x, v) -> (J, theta) can be found such that

Equation 6   (6)

The constants Ji, called the actions, define the radii of the various cross-sections of the torus while the angles thetai define the position on the torus (Figure 1). The dimensionality of a torus is therefore equal to that of an orbit in configuration space. Each point on an orbit maps to 2N points on its torus; for N = 2, these four points correspond to the four velocity vectors (± vx, ± vy) through a given configuration space point. The omegai are called fundamental frequencies; in the generic case, they are incommensurable, i.e. their ratios can not be expressed as ratios of integers. Orbits defined by incommensurable frequencies map onto the entire surface of the torus, filling it densely after a sufficiently long time. For certain orbits, a resonance between the fundamental frequencies occurs, i.e. n . omega = 0 with n = (l, m) a nonzero integer vector. In two degrees of freedom such a resonance implies omega1 / omega2 = |m/l| and the orbit is periodic, closing on itself after |m| revolutions in theta1 and | l| revolutions in theta2. When N > 2, each independent resonance condition reduces the dimensionality of the orbit by one, and N - 1 such conditions are required for closure. With the exception of special Hamiltonians like the spherical harmonic oscillator, resonant tori comprise a set of measure zero, although they are dense in the phase space.

Figure 1

Figure 1. Invariant torus defining the motion of a regular orbit in a two-dimensional potential. The torus is determined by the values of the actions J1 and J2; the position of the trajectory on the torus is defined by the angles theta1 and theta2, which increase linearly with time, thetai = omegait + theta0. If the torus is resonant, lomega1 + momega2 = 0, the orbit is periodic, returning to its starting point after |m| revolutions in theta1 and |l| revolutions in theta2.

In potentials that support N global integrals, like that of the Perfect Ellipsoid, all trajectories lie on tori and the Hamiltonian can be written

Equation 7   (7)

with

Equation 8   (8)

The tori of an integrable system are nested in a completely regular way throughout phase space. According to the KAM theorem, these tori will survive under small perturbations of H if their frequencies are sufficiently incommensurable ([Lichtenberg & Lieberman 1992]). Resonant tori may be strongly deformed even under small perturbations, however, leading to a complicated phase-space structure of interleaved regular and chaotic regions. Where tori persist, the motion can be characterized in terms of N local integrals. Where tori are destroyed, the motion is chaotic and the orbits move in a space of higher dimensionality than N.

While (local) action-angle variables are guaranteed to exist if the motion is regular, there is no general, analytic technique for calculating the (J, theta) from the (x, v). This is unfortunate since the representation of an orbit in terms of its action-angle variables is maximally compact, reducing the 2N - 1 variables on the energy surface to just N angles. In addition, time-averaged quantities become trivial to compute since the probability of finding a star on the torus is uniform in the angle variables. For instance, the configuration space density is

Equation 9   (9)

which can be trivially evaluated given x(theta). Finally, the actions are conserved under slow deformations of the potential, a useful property if one wishes to compute the evolution of a galaxy that is subject to some gradual perturbing force.

Fortunately, a number of algorithms have been developed in recent years for numerically extracting the relations between the Cartesian and action-angle variables. Most of this work has been directed toward 2 DOF systems, making it applicable to axisymmetric potentials or to motion in a principal plane of a nonrotating triaxial galaxy. Generalizations to three degrees of freedom are straightforward in principle, although in practice some of the algorithms described here become inefficient when N > 2.

2.1. Iterative Approaches

2.1.1. Perturbative Methods

In canonical perturbation theory, one begins by writing the Hamiltonian as the sum of two terms,

Equation 10   (10)

where H0 is integrable and epsilon is a (hopefully small) perturbation parameter. One then seeks a canonical transformation to new action-angle variables such that the Hamiltonian is independent of theta. The transformation is found by expanding the generating function in powers of epsilon and solving the Hamilton-Jacobi equation successively to each order. The KAM theorem states that such a transformation will sometimes exist, at least when epsilon is sufficiently small and when the unperturbed omegai are sufficiently far from commensurability. In a system with one degree of freedom, this approach is equivalent to Lindstedt's (1882) method in which both the amplitude and frequency of the oscillation are expanded as power series in epsilon; the result is a uniformly convergent series solution for the motion. Davoust (1983a, b, c) and Scuflaire (1995) presented applications of Lindstedt's method to periodic motion in simple galactic potentials.

In systems with two or more degrees of freedom, the transformation to action-angle variables is not usually expressible as a power series in epsilon. The reason is the "problem of small denominators": a Fourier expansion of H1 in terms of the thetai will generally contain terms with n . omega approx 0 which cause the perturbation series to blow up. This phenomenon is indicative of a real change in the structure of phase space near resonances: if the perturbation parameter epsilon is small, H0 + epsilon H1 can be topologically different from H0 only if H1 is large. However, a number of techniques have been developed to formally suppress the divergence. One approach, called secular perturbation theory, is useful near resonances in the unperturbed Hamiltonian, l omega1 + momega2 = 0. One first transforms to a frame that rotates with the resonant frequency; in this frame, the new angle variable measures the slow deviation from resonance. The Hamiltonian is then averaged over the other, fast angle variable. To lowest order in epsilon, the new action is lJ2 + mJ1 and the motion in the slow variable can be found by quadrature. Solutions obtained in this way are asymptotic approximations to the exact ones, i.e. they differ from them by at most O(epsilon) for times of order epsilon-1. Whether they are good descriptions of the actual motion depends on the strength of the perturbation and the timescale of interest.

Verhulst (1979) applied the averaging technique to motion in the meridional plane of an axisymmetric galaxy. He expanded the potential in a quartic polynomial about the circular orbit in the plane, thus restricting his results to epicyclic motion but at the same time guaranteeing a finite Fourier expansion for H1. He related the fundamental frequencies by setting omegaz2 = (r/s)2 omegax2 + delta(epsilon) and found solutions to first order in epsilon for various choices of the integers (r, s). For r = s = 1 he recovered Contopoulos's (1960) famous "third integral," as well as Saaf's (1968) formal integral for a quartic potential. De Zeeuw & Merritt (1983) and de Zeeuw (1985a) applied Verhulst's formalism to motion near the center of a triaxial galaxy with an analytic core. Here the unperturbed omegai are the frequencies of harmonic oscillation near the center. De Zeeuw & Merritt showed that the choice r = s = 1 produced a reasonable representation of the orbital structure in the plane of rotation, including the important 1 : 1 closed loop orbits that generate the three-dimensional tubes. Robe (1985, 1986, 1987) used an exact technique to recover the periodic orbits in the case of 1 : 1 : 1 resonance and investigated their stability.

Gerhard & Saha (1991) compared the usefulness of three perturbative techniques for reproducing the meridional-plane motion in axisymmetric models. They considered: (1) Verhulst's (1979) averaging technique; (2) a resonant method based on Lie transforms; and (3) a "superconvergent" method. The latter was proposed originally by Kolmogorov (1954) and developed by Arnold (1963) and Moser (1967) in their proof of the KAM theorem. The KAM method can only generate orbit families that are present in H0. Gerhard & Saha took for their unpertured potential the spherical isochrone and set Phi1 = g(r) P20(costheta). They were able to reproduce the main features of the motion using the resonant Lie transform method, including the box orbits that are not present when Phi1 = 0. The KAM method could only generate loop orbits but did so with great accuracy when taken to high order. Dehnen & Gerhard (1993) used the approximate integrals generated by the resonant Lie-transform method to construct three-integral models of axisymmetric galaxies, as described in Section 3.3.

2.1.2. Nonperturbative Methods

The failure of perturbation expansions to converge reflects the change in phase space topology that occurs near resonances in the unperturbed Hamiltonian. As an alternative to perturbation theory, one can assume that a given orbit is confined to a torus and solve directly for the action-angle variables in terms of x and v. If the orbit is indeed regular, such a solution is guaranteed to exist and corresponds to a perturbation expansion carried out to infinite order (assuming the latter is convergent). For a chaotic orbit, the motion is not confined to a torus but one might still hope to derive an approximate torus that represents the average behavior of a weakly chaotic trajectory over some limited interval of time.

Ratcliff, Chang & Schwarzschild (1984) pioneered this approach in the context of galactic dynamics. They noted that the equations of motion of a 2D orbit,

Equation 11   (11)

can be written in the form

Equation 12   (12)

where theta1 = omega1t and theta2 = omega2t, the coordinates on the torus. If one specifies omega1 and omega2 and treats ðPhi / ðx and ðPhi / ðy as functions of the thetai, equations (12) can be viewed as nonlinear equations for x(theta1, theta2) and y(theta1, theta2). No very general method of solution exists for such equations; iteration is required and success depends on the stability of the iterative scheme. Ratcliff et al. chose to express the coordinates as Fourier series in the angle variables,

Equation 13   (13)

Substituting (13) into (12) gives

Equation 14   (14)

where the right hand side is again understood to be a function of the angles. Ratcliff et al. truncated the Fourier series after a finite number of terms and required equations (14) to be satisfied on a grid of points around the torus. They then solved for the Xn by iterating from some initial guess. Convergence was found to be possible if the initial guess was close to the exact solution. Guerra & Ratcliff (1990) applied a similar algorithm to orbits in the plane of rotation of a nonaxisymmetric potential.

Another iterative approach to torus reconstruction was first developed in the context of semiclassical quantum theory by Chapman, Garrett & Miller (1976). Binney and collaborators ([McGill & Binney 1990]; [Binney & Kumar 1993]; [Kaasalainen & Binney 1994a], 1994b; [Kaasalainen 1994], 1995a, b) further developed the technique and applied it to galactic potentials. One starts by dividing the Hamiltonian H into separable and non-separable parts H0 and H1, as in equation (10); however epsilonH1 is no longer required to be small. One then seeks a generating function S that maps the known tori of H0 into tori of H. For a generating function of the F2-type (Goldstein 1980), one has

Equation 15   (15)

where (J, theta) and (J', theta') are the action-angle variables of H0 and H respectively. The generator S is determined, for a specified J', by substituting the first of equations (15) into the Hamiltonian H(J, theta) and requiring the result to be independent of theta. One then arrives at H(J'). Chapman et al. showed that a sufficiently general form for S is

Equation 16   (16)

where the first term is the identity transformation, and they evaluated a number of iterative schemes for finding the Sn. One such scheme was found to recover the results of first-order perturbation theory after a single iteration.

The generating function approach is useful for assigning energies to actions, H(J'), but most of the other quantities of interest to galactic dynamicists require additional effort. For instance, equation (15) gives theta'(theta) as a derivative of S, but since S must be computed separately for every J' its derivative is likely to be ill-conditioned. Binney & Kumar (1993) and Kaasalainen & Binney (1994a) discussed two schemes for finding theta'(theta); the first required the solution of a formally infinite set of equations, while the latter required multiple integrations of the equations of motion for each torus.

Another feature of the generating function approach is its lack of robustness. Kaasalainen & Binney (1994a) noted that the success of the method depends somewhat on the choice of H0. For box orbits, which are most naturally described as coupled rectilinear oscillators, they found that a harmonic-oscillator H0 gave poor results unless an additional point transformation was used to deform the rectangular orbits of H0 into narrow-waisted boxes like those in typical galactic potentials. Kaasalainen (1995a) considered orbits belonging to higher-order resonant families and found that it was generally necessary to define a new coordinate transformation for each family.

Warnock (1991) presented a modification of the Chapman et al. algorithm in which the generating funtion S was derived by numerically integrating an orbit from appropriate initial conditions, transforming the coordinates to (J, theta) of H0 and interpolating J on a regular grid in theta. The values of the Sn then follow from the first equation of (15) after a discrete Fourier transform. Kaasalainen & Binney (1994b) found that Warnock's scheme could be used to substantially refine the solutions found via their iterative scheme.

Having computed the energy on a grid of J' values, one can interpolate to obtain the full Hamiltonian H(J'1, J'2). If the system is not in fact completely integrable, this H may be rigorously interpreted as smooth approximation to the true H ([Warnock & Ruth 1991], 1992) and can be taken as the starting point for secular perturbation theory. Kaasalainen (1994, 1995b) developed this idea and showed how to recover accurate surfaces of section in the neighborhood of low-order resonances in the planar logarithmic potential.

Percival (1977) described a variational principle for constructing tori. His technique has apparently not been implemented in the context of galactic dynamics.

2.2. Trajectory-Following Approaches

A robust and powerful alternative to the generating function approach is to construct tori by Fourier decomposition of the trajectories. Trajectory-following algorithms are based on the fact that integrable motion is quasiperiodic; in other words, in any canonical coordinates (p, q), the motion can be expressed as

Equation 17   (17)

where the omegai are the N fundamental frequencies on the torus. It follows from equations (17) that the Fourier transform of x(t) or v(t) will consist of a set of spikes at discrete frequencies omegak that are linear combinations of the fundamental frequencies. An analysis of the frequency spectrum yields both the fundamental frequencies and the integer vectors n associated with each spike. The relation between the Cartesian coordinates and the angles follows immediately from equation (17). For instance, the x coordinate in a 2 DOF system becomes

Equation 18   (18)

and similarly for y(t). The actions can be computed from Percival's (1974) formulae,

Equation 19   (19)

thus yielding the complete map (x,v) -> (J, theta). Binney & Spergel (1982) pioneered the trajectory-following approach in galactic dynamics, using a least-squares algorithm to compute the Xk. They were able to recover the fundamental frequencies in a 2 DOF potential with a modest accuracy of ~ 0.1% after ~ 25 orbital periods. Binney & Spergel (1984) used Percival's formula to construct the action map for orbits in a principal plane of the logarithmic potential.

A major advance in trajectory-following algorithms was made by Laskar (1988, 1990), who developed a set of tools, the "numerical analysis of fundamental frequencies" (NAFF), for extracting the frequency spectra of quasiperiodic systems with very high precision. The NAFF algorithm consists of the following steps:

  1. Integrate an orbit for a time T and record the phase space variables at M equally spaced intervals. Translate each time series f (t) to an interval [- T/2, T/2] symmetric about the time origin.

  2. Using a discrete Fourier transform, construct an approximation to the frequency spectrum of f (t) and identify the peaks. The location of any peak omegak will be defined to an accuracy of ~ 1/MT.

  3. Refine the estimate of the location of the strongest peak by finding the maximum of the function

    Equation 20   (20)

    where chi(t) = 1 + cos(2pit/T) is the Hanning window function. The integral can be approximated by interpolating the discretely-sampled f (t). The Hanning filter broadens the peak but greatly reduces the sidelobes, allowing a very precise determination of omegak.

  4. Compute the amplitude Xk by projecting eiomegakt onto f (t), and subtract this component from the time series.

  5. Repeat steps 3 and 4 until the residual function does not significantly decrease following subtraction of another term. Since subsequent components eiomegajt will not be mutually orthogonal, a Gram-Schmidt procedure is used to construct orthonormal basis functions before carrying out the projection in step 4.

  6. Identify the integer vector nk = (lk, mk) associated with each omegak.

Laskar's algorithm recovers the fundamental frequencies with an error that falls off as T-4 ([Laskar 1996]), compared with ~ T-1 in algorithms like Binney & Spergel's (1982). Even for modest integration times of ~ 102 orbital periods, the NAFF algorithm is able to recover fundamental frequencies with an accuracy of ~ 10-8 or better. Such extraordinary precision allows the extraction of a large number of components from the frequency spectrum, hence a very precise representation of the torus.

Papaphilippou & Laskar (1996) applied the NAFF algorithm to 2DOF motion in a principal plane of the logarithmic potential. They experimented with different choices for the quantity f whose time series is used to compute the frequency spectrum. Ideally, one would choose f to be an angle variable, in terms of which the frequency spectrum reduces to a single peak, but the angles are not known a priori and the best one can do is to use the angle variables corresponding to some well-chosen Hamiltonian. However Papaphilippou & Laskar found that the convergence of the quasiperiodic expansion was only weakly affected by the choice of f; for most orbits, Cartesian coordinates (or polar coordinates in the case of loop orbits) were found to work almost as well as other choices. This result implies that trajectory-following methods are more easily automated than generating function methods which require a considerable degree of cleverness in the choice of coordinates.

Papaphilippou & Laskar (1996) focussed on the fundamental frequencies rather than the actions for their characterization of the tori, in part because the frequencies can be obtained with more precision than the actions, but also because KAM theory predicts that the structure of phase space is determined in large part by resonances between the omegai. They defined the "frequency map," the curve of omega2 / omega1 values determined by a set of orbits of a given energy; this curve is discontinuous whenever the initial conditions pass over a resonance associated with chaotic motion. The important resonances, and the sizes of their associated chaotic regions, are immediately apparent from the frequency map. Papaphilippou & Laskar showed that most of the chaos in the logarithmic potential was associated with the unstable short-axis orbit, a 1 : 1 resonance, but they were also able to identify significant chaotic zones associated with higher-order resonances like the 1 : 2 banana orbit.

One shortcoming of trajectory-following algorithms is that they must integrate long enough for the orbit to adequately sample its torus. When the torus is nearly resonant, n . omega approx 0, the orbit is restricted for long periods of time to a subset of the torus and the required integration interval increases by a factor ~ 1/n . omega. Another problem is the need to sample the time series with very high frequency in order to minimize the effects of aliasing.

Next Contents Previous