1: \documentclass[useAMS,usenatbib]{mn2e}
2: \usepackage{graphicx}
3: \usepackage{pslatex}
4: \usepackage{natbib}
5: \usepackage{txfonts}
6: \newcommand{\Eqref}[1]{(\ref{#1})}
7: \newcommand{\sfrac}[2]{\mbox{$\frac{#1}{#2}$}}
8: \newcommand{\roms}[1]{{\mbox{\rm #1}}}
9: \newcommand{\sgn}[1]{\textrm{sgn}(#1)}
10: \newcommand{\cH}{\mathcal{H}}
11: \newcommand{\cK}{\mathcal{K}}
12: \newcommand{\cM}{\mathcal{M}}
13: %
14: \newcommand\Chi{{(\chi^2_\nu)^{1/2}}}
15: \def\deg{{\rm o}}
16: \def\idm#1{{\mbox{\scriptsize #1}}}
17: \renewcommand{\vec}[1]{\mathbf{#1}}
18:
19: \voffset -0.5in
20:
21: \begin{document}
22:
23: \title{The long-term stability of extrasolar system HD~37124.
24: Numerical study of resonance effects.
25: }
26: \author[Go\'zdziewski, Breiter and Borczyk]
27: {Krzysztof Go\'zdziewski$^1$\thanks{Toru\'n Centre for Astronomy,
28: N. Copernicus University, Poland, k.gozdziewski@astri.uni.torun.pl},
29: S\l awomir Breiter$^2$\thanks{Astronomical Observatory of A. Mickiewicz University,
30: S\l oneczna 36, PL 60-286 Pozna\'n, Poland, breiter@amu.edu.pl},
31: Wojciech Borczyk$^2$\thanks{Astronomical Observatory of A. Mickiewicz University,
32: S\l oneczna 36, PL 60-286 Pozna\'n, Poland, bori@moon.astro.amu.edu.pl}
33: }
34: \maketitle
35: \begin{abstract}
36: We describe numerical tools for the stability analysis of extrasolar
37: planetary systems. In particular, we consider the relative Poincar\'e variables
38: and symplectic integration of the equations of motion. We apply the tangent map
39: to derive a numerically efficient algorithm of the fast indicator MEGNO (a
40: measure of the maximal Lyapunov exponent) that helps to distinguish chaotic
41: and regular configurations. The results concerning the three-planet extrasolar
42: system HD~37124 are presented and discussed. The best fit solutions found in
43: earlier works are studied more closely. The system involves Jovian planets with
44: similar masses. The orbits have moderate eccentricities, nevertheless the best
45: fit solutions are found in dynamically active region of the phase space. The
46: long term stability of the system is determined by a net of low-order two-body
47: and three-body mean motion resonances. In particular, the three-body resonances
48: may induce strong chaos that leads to self-destruction of the system after Myrs
49: of apparently stable and bounded evolution. In such a case, numerically
50: efficient dynamical maps are useful to resolve the fine structure of the phase
51: space and to identify the sources of unstable behavior.
52: \end{abstract}
53: \begin{keywords}
54: {extrasolar planets---Doppler technique---stars:individual HD~37124---N-body
55: problem---numerical methods
56: }
57: \end{keywords}
58:
59: \section{Introduction}
60:
61: {Understanding
62: }
63: the extrasolar planetary systems has became a major challenge for contemporary
64: astronomy. One of the most difficult problems in this field concerns the
65: orbital stability of such systems. Usually, the investigations of long-term
66: evolution are the domain of direct, numerical integrations. The stability of
67: extrasolar systems is often understood in terms of the Lagrange definition
68: implying that orbits remain well bounded over an arbitrarily long time. Other
69: definitions may be formulated as well, like the astronomical stability
70: \citep{Lissauer1999} requiring that the system persists over a very long, Gyr
71: time-scale, or Hill stability \citep{Szebehely1984} that requires the constant
72: ordering of the planets. In our studies, we prefer a more formal and stringent
73: approach related to the fundamental Kolmogorov-Arnold-Theorem (KAM), see
74: \cite{Arnold1978}. Planetary systems, involving a dominant mass of the parent
75: star and significantly smaller planetary masses, are well modeled by
76: close-to-integrable, Hamiltonian dynamical systems. It is well known, that their
77: evolution may be quasi-periodic (with a discrete number of fundamental
78: frequencies, forever stable), periodic (or resonant; stable or unstable) or
79: chaotic (with a continuous spectrum of frequencies, and unstable). In the last
80: case, initially close phase trajectories diverge exponentially, i.e., their
81: Maximum Lyapunov Characteristic Exponent (MLCE, denoted also with $\sigma$) is
82: positive. In general, the distinction between regular and chaotic trajectories
83: is a very difficult task that may be resolved only with numerical methods
84: relying on efficient and accurate integrators of the equations of motion.
85:
86: The main motivation of this paper is to describe numerical tools that are useful
87: for studies of the dynamical stability and to apply them to the HD~37124 system
88: \citep{Vogt2005}. We recall the fundamentals of relative canonical Poincar\'e
89: variables as -- in our opinion -- one of the best frameworks for symplectic
90: integrators. These {canonical} variables are well suited for the
91: construction of a \citet{LasRob} composition method that improves a classical
92: Wisdom-Holman (W-H) algorithm \citep{WH:91}. We supplement the integrator with a
93: propagator of the associated symplectic tangent map that approximates the
94: solution of variational equations \citep{MikIn:99}. Finally, we compare two fast
95: indicators that reveal the character of phase trajectories. The first one is a
96: relatively simple method for resolving fundamental frequencies and spectral
97: properties of a close-to-integrable Hamiltonian system -- a so called Spectral
98: Number (SN), invented by \citet{Michtchenko2001}. The second indicator belongs
99: to the realm of the Lyapunov exponent based algorithms; we chose the numerical
100: tool developed by \cite{Cincotta2000,Cincotta2003} under the name of MEGNO. In
101: this work, we refine the algorithm of MEGNO that makes explicit use of the
102: symplectic tangent map \citep{Gozdziewski2003b}.
103:
104: As a non-trivial application of the presented numerical tools, we consider the
105: 3-planet system hosted by the HD~37124~star \citep{Vogt2005}. It has been
106: discovered by the radial velocity (RV) technique. The recent model of the RV
107: observations of HD~37124 predicts three equal Jovian type planets with masses
108: $\sim 0.6$~m$_{\idm{J}}$ in orbits with moderate eccentricities. In such a case,
109: the application of symplectic integrators without regularization is particularly
110: advantageous thanks to the numerical efficiency (long time-steps) and accuracy
111: (the total energy does not have a secular error and the angular momentum
112: integral is conserved). The number of multi-planet systems resembling the
113: architecture of the Solar system increases\footnote{For a recent statistics of
114: the discoveries, see Jean Schneider's Extrasolar Planets Encyclopedia,
115: http://exoplanets.eu.}. Hence, our approach may be useful in other cases.
116:
117: \section{Numerical tools}
118:
119: According to the classical results of celestial mechanics, the $N$-body problem
120: has only 10 integrals of motion for all $N > 2$; they consist of 6 integrals of
121: barycenter, 3 integrals of angular momentum and the energy integral. The
122: integrals of barycenter play a very particular role in the studies of an
123: $N$-body system dynamics. First, they define the origin of an inertial reference
124: frame in terms of the mutual distances and velocities of the bodies considered,
125: thus dismissing the need of some extrinsic absolute frame. But what is more
126: important, being linear forms of coordinates and momenta they allow a unique
127: reduction of the system, lowering the number of degrees of freedom by three,
128: with no loss of information. This is why we can solve the relative two-body
129: problem and then recover the motion of both masses with respect to their center
130: of mass. And this is why we can approximately solve the heliocentric motion of
131: planets, recovering the barycentric evolution \textit{a posteriori}.
132:
133: Within the framework of Hamiltonian mechanics, the reduction is usually achieved
134: by means of a transformation to one of the two common variable types: relative
135: Jacobi variables, or ''heliocentric'' Poincar\'e variables
136: \citep[Ch.~XIII]{Whittaker}. We focus on the latter set, because it offers the
137: best choice in many aspects. We introduce the basic ideas related to the
138: Poincar\'e variables and we derive them as a Mathieu transformation; this way is
139: simpler and more intuitive than the procedure based upon a generating function
140: that was presented by \citet[p.~343]{Whittaker} or \citet{DLL:98}.
141: {
142: Then we discuss the setup of the Hamiltonian within the framework
143: of Wisdom-Holman type integrators.
144: }
145:
146: \subsection{Poincar\'e variables basics}
147:
148: Let us consider a system consisting of $N+1$ material points with masses $m_0,
149: \ldots ,m_N$. We define a barycentric position vector $\vec{p} \in
150: \mathbb{R}^{3N}$ and its canonical conjugate momentum $\vec{P} \in
151: \mathbb{R}^{3N}$ as
152: \begin{equation}
153: \label{pP}
154: \vec{p}=\left[ \begin{array}{c}
155: \vec{p}_0 \\
156: \vec{p}_1 \\
157: \vdots \\
158: \vec{p}_N \\
159: \end{array}
160: \right], \qquad
161: \vec{P}=\left[ \begin{array}{c}
162: \vec{P}_0 \\
163: \vec{P}_1 \\
164: \vdots \\
165: \vec{P}_N \\
166: \end{array}
167: \right].
168: \end{equation}
169: Then, the barycentric equations
170: of motion can be derived from the Hamiltonian function
171: \citep{Laskar:90}
172: \begin{equation}\label{Hbar}
173: \cH(\vec{p},\vec{P}) =
174: \sum_{i=0}^N \frac{\vec{P}_i^2}{2\,m_i}
175: - \sum_{i=0}^{N}
176: \sum_{ j=i+1}^N
177: \frac{k^2\,m_i\,m_j}{\Delta_{ij}} ,
178: \end{equation}
179: where the mutual distance $\Delta_{ij}$ is
180: \begin{equation}\label{Delta}
181: \Delta_{ij} = ||\vec{p}_j - \vec{p}_i||,
182: \end{equation}
183: and $k$ stands for the Gaussian gravity constant.
184:
185: The Hamiltonian \Eqref{Hbar} admits six integrals of barycenter
186: \begin{equation} \label{intbar}
187: \sum_{i=0}^N m_i\,\vec{p}_i = \vec{0}, \qquad
188: \sum_{i=0}^N \vec{P}_i = \vec{0},
189: \end{equation}
190: as well as the energy integral $\cH = \mbox{const}$, and
191: three angular momentum integrals
192: \begin{equation}\label{Gint}
193: \vec{G} = \sum_{i=0}^N \vec{p}_i \times \vec{P}_i = \mathbf{const}.
194: \end{equation}
195: The integrals are usually exploited as the accuracy control tool
196: when the differential equations of motion
197: \begin{equation}\label{Beq}
198: \dot{ \vec{p}} = \frac{\partial \cH}{\partial \vec{P}},
199: \qquad \dot{\vec{P}} = - \frac{\partial \cH}{\partial \vec{p}},
200: \end{equation}
201: are solved numerically.
202:
203: Instead of solving the $6N+6$ order system \Eqref{Beq}, it is often desirable to
204: study the relative motion of $N$ bodies with respect to the material point
205: $m_0$. But if the integration method to be applied is symplectic, it is
206: necessary to use the Hamiltonian equations of motion, hence the necessity of
207: defining a canonical transformation $(\vec{p},\vec{P},\cH) \leftrightarrows
208: (\vec{r},\vec{R},\cK)$ with new, relative coordinates $\vec{r}$ and momenta
209: $\vec{R}$.
210:
211: A naive, straightforward approach would consist in postulating $ \vec{r}_i =
212: \vec{p}_i - \vec{p}_0$ for all $i$. This leads to $\vec{r}_0 = \vec{0}$ and the
213: Jacobian matrix of the transformation becomes singular; such a transformation
214: cannot be canonical regardless of the choice of the momenta. However, the
215: difficulty can be easily circumvented if we change the definition of
216: $\vec{r}_0$. \citet{Poincare:96} proposed
217: \begin{equation}
218: \label{pt}
219: \begin{array}{rcl}
220: \vec{r}_i &=& \vec{p}_i - \vec{p}_0, \qquad \mathrm{for~}i=1,\ldots,N, \\
221: \vec{r}_0 &=& \vec{p}_0, \\
222: \end{array}
223: \end{equation}
224: whereas \citet{DLL:98} chose $\vec{r}_0$ as the position of the barycenter in
225: some arbitrary inertial frame. For the barycentric system, the latter choice
226: amounts to a ''differentiable zero'' $\vec{r}_0 = \left( \sum_{i=0}^N m_i
227: \right)^{-1} \sum_{i=0}^N m_i \vec{p}_i = \vec{0}$. It turns out, that both
228: starting points lead to the same result, so we continue our presentation
229: assuming the Poincar\'e choice \Eqref{pt}.
230:
231: Retaining $\vec{r}_0$ as the position of the reference body $m_0$ with respect
232: to the barycenter, we can perform a canonical extension of the time-independent
233: point transformation \Eqref{pt},
234: {
235: requesting the
236: Mathieu transformation condition
237: }
238: \citep[p.~301]{Whittaker}
239: \begin{equation}\label{canext}
240: \vec{P} \cdot \mathrm{d}\vec{p} = \vec{R} \cdot \mathrm{d}\vec{r},
241: \end{equation}
242: or, explicitly,
243: \begin{equation}\label{canext1}
244: \sum_{i=0} \vec{P}_i \cdot \mathrm{d}\vec{p}_i =
245: \vec{R}_0 \cdot \mathrm{d}\vec{p}_0
246: + \sum_{i=1}^N \vec{R}_i \cdot
247: \left( \mathrm{d}\vec{p}_i - \mathrm{d}\vec{p}_0\right) .
248: \end{equation}
249: Equating the coefficient
250: of each differential $\mathrm{d}\vec{p}_i$ to zero, we find the
251: new momenta
252: \begin{equation}
253: \label{ptm}
254: \begin{array}{rcl}
255: \vec{R}_i &=& \vec{P}_i, \qquad \mbox{for} \qquad i=1,\ldots,N, \\
256: \vec{R}_0 &=& \sum_{i=0}^N \vec{P}_i ~=~\vec{0}. \\
257: \end{array}
258: \end{equation}
259: The fact, that $\vec{R}_0=\vec{0}$ is a direct consequence of the
260: integrals of barycenter \Eqref{intbar}.
261:
262: Equations \Eqref{pt} and \Eqref{ptm}, that are due to \citet{Poincare:96},
263: define the canonical relative variables. The coordinates $\vec{r}$ consist of
264: the positions with respect to the reference body $m_0$, save for the $\vec{r}_0$
265: that is measured with respect to the barycenter. The momenta $\vec{R}$ are
266: measured with respect to the barycenter, save for $\vec{R}_0$ that can be
267: understood as measured with respect to $m_0$ and hence it is zero (although with
268: nonvanishing partials with respect to $\vec{P}_i$).
269:
270: The most important feature is that the new Hamiltonian $\cK$, obtained by the
271: simple substitution of the transformation equations into $\cH$, does not depend
272: neither on the coordinates, nor on the momenta of $m_0$. Indeed, one obtains
273: \citep{Poincare:96,Whittaker,Hagihara:70,Deprit:83,Laskar:90,Laskar:91}
274: \begin{eqnarray}
275: \cK & = & \frac{1}{2} \sum_{i=1}^{N} \left( \frac{1}{m_0}
276: + \frac{1}{m_i} \right)
277: \vec{R}_i^2 + \frac{1}{m_0}
278: \sum_{i=1}^{N} \sum_{j=i+1}^{N}
279: \vec{R}_i \cdot \vec{R}_j - \nonumber \\
280: & & -
281: \sum_{i=1}^{N} \frac{k^2 m_0 m_i}{r_i} -
282: \sum_{i=1}^{N} \sum_{j=i+1}^{N} \frac{k^2 m_i m_j}{\Delta_{ij}}, \label{Nc:hamP}
283: \end{eqnarray}
284: where $\Delta_{ij}=||\vec{p}_j-\vec{p}_i|| = ||\vec{r}_j-\vec{r}_i||$,
285: or, equivalently \citep{DLL:98,Chambers:99}
286: \begin{eqnarray}
287: \cK & = & \frac{1}{2} \sum_{i=1}^{N}
288: \frac{1}{m_i}
289: \vec{R}_i^2 + \frac{1}{2m_0}\left(
290: \sum_{i=1}^{N}
291: \vec{R}_i \right)^2 - \nonumber \\
292: & & -
293: \sum_{i=1}^{N} \frac{k^2 m_0 m_i}{r_i} -
294: \sum_{i=1}^{N} \sum_{j=i+1}^{N} \frac{k^2 m_i m_j}{\Delta_{ij}}. \label{HDLL}
295: \end{eqnarray}
296:
297: An important fact to be remembered is that the Hamiltonian $\cK$ has the form
298: \Eqref{Nc:hamP} or \Eqref{HDLL} only if the substitution of the barycenter
299: integrals \Eqref{intbar} has been performed. Thus it cannot serve to obtain the
300: equations for $\dot{\vec{r}}_0$ or $\dot{\vec{R}}_0$. However, once we know all
301: the remaining $\vec{r}_i$ and $\vec{R}_i$, the $\vec{r}_0$ values can be easily
302: computed from Equations \Eqref{intbar} and \Eqref{pt}, whereas
303: $\vec{R}_0=\vec{0}$ by the definition. For all the remaining bodies
304: \begin{equation}\label{Peq}
305: \dot{\vec{r}}_i = \frac{\partial \cK}{\partial \vec{R}_i},
306: \qquad \dot{\vec{R}}_i = - \frac{\partial \cK}{\partial \vec{r}_i},
307: \end{equation}
308: and we will assume that $i=1,\ldots,N$ throughout the rest of this
309: paper.
310:
311: A remarkable property of the Poincar\'e variables is
312: \begin{equation}\label{GP}
313: \vec{G} = \sum_{i=0}^N \vec{p}_i \times \vec{P}_i =
314: \sum_{i=1}^N \vec{r}_i \times \vec{R}_i,
315: \end{equation}
316: which means, that the total angular momentum of the reduced
317: system of $N$ bodies evaluated by means of the
318: Poincar\'e variables is the same as the angular momentum
319: of $N+1$ bodies evaluated in the barycentric frame.
320:
321:
322: If the reference body mass $m_0 \gg m_i$, the Hamiltonian $\cK$ can be easily
323: partitioned into the unperturbed, Keplerian part and a small perturbation
324: proportional to the greatest of $m_i$. From the point of view of analytical
325: theories using these variables \citep{YuasaHori:79,Hori:85}, it is preferable
326: to split $\cK$ into the sum
327: \begin{equation}\label{parta}
328: \cK = \cK_0^{(a)} + \cK_1^{(a)},
329: \end{equation}
330: where \citep{Laskar:90,Laskar:91}
331: \begin{eqnarray}
332: \cK_0^{(a)} & = & \frac{1}{2} \sum_{i=1}^{N}
333: \frac{m_0+m_i}{m_0\,m_i} \vec{R}_i^2 -
334: \sum_{i=1}^{N} \frac{k^2 m_0 m_i}{r_i},
335: \label{Nc:K0}\\
336: \cK_1^{(a)} & = & \frac{1}{m_0}
337: \sum_{i=1}^{N} \sum_{j=i+1}^{N}
338: \vec{R}_i \cdot \vec{R}_j -
339: \sum_{i=1}^{N} \sum_{j=i+1}^{N} \frac{k^2 m_i m_j}{\Delta_{ij}} .
340: \label{Nc:K1}
341: \end{eqnarray}
342: The principal part $\cK_0^{(a)}$ defines $N$ \emph{relative} two-body problems.
343: The perturbation $\cK_1^{(a)}$ involves not only the mutual interactions between
344: the minor bodies $m_i$, but also the momenta related terms that replace the
345: usual ''indirect part'' of the perturbing function present in noncanonical
346: relative $(N+1)$-body problem \citep{Poincare:05}. It is due to this term, that
347: the Poincar\'e variables were considered somehow handicapped; the objection that
348: velocities are no longer tangent to the momenta became almost a proverb,
349: although many non-inertial reference frames have the same property, the
350: restricted three-body problem being the best example. This objection has
351: fortunately ceased to be taken seriously; for example, \citet{FMM:04}
352: successfully use orbital elements evaluated from the Poincar\'e momenta
353: $\vec{R}$. In this paper we will use alternatively two types of orbital
354: elements. The \textit{osculating elements} are computed by the usual two body
355: formulae from astrocentric positions $\vec{r}_i$ and velocities
356: $\dot{\vec{r}_i}$; we use them in all plot labels and orbital data tables. But
357: in the calculation of spectral numbers or in the definitions of resonance
358: arguments, we use orbital elements computed from $\vec{r}_i$ and
359: $m_0m_i/(m_0+m_i)\,\vec{R}_i$, calling them \textit{contact elements} after
360: \citet{Brumberg91}.
361: [{In fact, the transformation between astrocentric positions -- barcentric momenta
362: and the contact elements can be still done with the usual two-body formulae
363: \citep{Morbidelli:book}}]. The former are commonly used in literature, whereas
364: the latter offer a better behavior from the dynamical point of view.
365: {
366: The superiority of contact elements results from the fact, that the reference frame
367: of Cartesian momenta is inertial, hence the influence of noninertial forces is
368: reduced to purely kinematical contribution. The inferred Keplerian angles
369: ($\omega,\Omega, \mathcal{M}$) and the conjugate momenta can be interpreted as
370: canonical Delaunay's elements \citep{Morbidelli:book}, so the derivation and
371: interpreation of the fundamental frequencies is straightforward.
372: }
373:
374: \subsection{Symplectic integration: two is a company}
375:
376: With the advent of symplectic integrators based on the Wisdom-Holman approach
377: \citep{WH:91}, the Poincar\'e variables became an attractive framework for the
378: numerical studies of planetary systems \citep{DLL:98,Chambers:99}. First of all,
379: similarly to the Jacobian coordinates, they reduce
380: {the number of
381: equations of motion} by 6 with respect to the barycentric problem. Thanks to the
382: possibility of splitting $\cK$ into the main part and a perturbation, they also
383: allow the construction of a symplectic integrator with the local truncation
384: error proportional to the product of $m_i/m_0$, hence a larger integration step
385: $h$ can be applied.
386:
387: Given a Hamiltonian $\mathcal{M} = \mathcal{M}_0 + \varepsilon \mathcal{M}_1$
388: with a small parameter $\varepsilon$, the W-H integrator is based on the
389: alternating application of maps $\Phi_{0,\tau}(\vec{r},\vec{R})$ and
390: $\Phi_{1,\tau}(\vec{r},\vec{R})$ that represent the solutions of equations of
391: motion derived from $\mathcal{M}_0$ and $\varepsilon \mathcal{M}_1$ alone, on
392: the interval from $t_0$ to $t_0+\tau $.
393: {
394: Moreover, each of the Hamiltonian parts should admit (a possibly simple)
395: analytical solution of the equations of motion.
396: }
397:
398: In the numerical applications, Hamiltonian $\cK$ is typically split differently
399: than in Equation \Eqref{parta}, namely \citep{DLL:98,Chambers:99}
400: \begin{equation}\label{partnew}
401: \cK = \cK_0 + \cK_1,
402: \end{equation}
403: where
404: \begin{eqnarray}
405: \cK_0 & = & \frac{1}{2} \sum_{i=1}^{N}
406: \frac{1}{m_i} \vec{R}_i^2 -
407: \sum_{i=1}^{N} \frac{k^2 m_0 m_i}{r_i},
408: \label{K0n}\\
409: \cK_1 & = & \frac{1}{2 m_0} \left( \sum_{i=1}^{N}
410: \vec{R}_i\right)^2 -
411: \sum_{i=1}^{N} \sum_{j=i+1}^{N} \frac{k^2 m_i m_j}{\Delta_{ij}} .
412: \label{K1n}
413: \end{eqnarray}
414: The unperturbed part $\cK_0$ has now a different meaning: it
415: still leads to $N$ relative two-body problems
416: \begin{eqnarray}\label{kep:1}
417: \dot{\vec{r}}_i & = & \frac{\partial \cK_0}{\partial \vec{R}_i} ~=~
418: \frac{\vec{R}_i}{m_i}, \\
419: \dot{\vec{R}}_i & = & - \frac{\partial \cK_0}{\partial \vec{r}_i}
420: ~=~ - \frac{k^2 m_0
421: m_i}{r_i^3}\,\vec{r}_i, \label{kep:2}
422: \end{eqnarray}
423: but this time they are the \emph{restricted} two-body problems
424: { with negligible masses $m_i$ or a fixed center of gravity:}
425: \begin{equation}\label{kep:12}
426: \vec{\ddot{r}}_i = - \frac{k^2 m_0}{r_i^3}\,\vec{r}_i.
427: \end{equation}
428:
429: The perturbation part $\cK_1$ remains proportional to $m_i/m_0$
430: and it is still a function of both coordinates $\vec{r}$ and momenta
431: $\vec{R}$. \citet{DLL:98} and then \citet{Chambers:99} considered it an
432: obstacle, so they further split $\cK_1$ into
433: \begin{equation}\label{K1sp}
434: \cK_1 = \cK_{11}(\vec{R}) + \cK_{12}(\vec{r}),
435: \end{equation}
436: obtaining elementary ''kick'' maps $\Phi_{11,\tau}$
437: \begin{eqnarray}
438: \vec{r}'_i & = & \vec{r}_i + \frac{\tau}{m_0} \sum_{j=1}^{N} \vec{R}_j, \\
439: \vec{R}'_i & = & \vec{R}_i .
440: \end{eqnarray}
441: and $\Phi_{12,\tau}$
442: \begin{eqnarray}
443: \vec{r}'_i & = & \vec{r}_i , \\
444: \vec{R}'_i & = &
445: \vec{R}_i - \tau \,\sum_{j=1, j\neq i}^{N}
446: \frac{k^2 m_i m_j}{\Delta_{ij}^3} \, \left( \vec{r}_i - \vec{r}_j \right).
447: \end{eqnarray}
448: In the formulas of both maps we add a prime to the symbols standing for the values of
449: coordinates and momenta at $t_0+\tau$, whereas unprimed symbols refer to the values at $t_0$.
450:
451: As the effect of the partition \Eqref{K1sp}, the classical ''leapfrog''
452: \begin{equation}
453: \label{leapr}
454: \Phi_\tau \approx \Phi_{1,\tau/2} \circ \Phi_{0,\tau} \circ \Phi_{1,\tau/2},
455: \end{equation}
456: was replaced by
457: \begin{equation}
458: \label{leapinc}
459: \Phi_\tau \approx \Phi_{11,\tau/2} \circ \Phi_{12,\tau/2} \circ \Phi_{0,\tau} \circ \Phi_{12,\tau/2}
460: \circ \Phi_{11,\tau/2}.
461: \end{equation}
462: According to \citet{DLL:98}, the ordering of $\Phi_{11}$ and $\Phi_{12,\tau}$ is
463: insignificant,
464: and, indeed, \citet{Chambers:99} interchanged them, using
465: \begin{equation}
466: \label{leapch}
467: \Phi_\tau \approx \Phi_{12,\tau/2} \circ \Phi_{11,\tau/2} \circ \Phi_{0,\tau} \circ \Phi_{11,\tau/2}
468: \circ \Phi_{12,\tau/2}.
469: \end{equation}
470: The interchange of the maps is justified by the fact that $\cK_{11}$ and $\cK_{12}$ commute, i.e.
471: the Poisson bracket $\{\cK_{11};\,\cK_{12}\}=0.$
472: In these circumstances
473: \begin{equation}
474: \label{compo}
475: \Phi_{1,\tau} = \Phi_{12,\tau} \circ \Phi_{11,\tau} = \Phi_{11,\tau }
476: \circ \Phi_{12,\tau},
477: \end{equation}
478: and we can concatenate both maps obtaining $\Phi_{1,\tau}$ in a compact form.
479: \begin{eqnarray}
480: \label{fi1a}
481: \vec{r}'_i &=& \vec{r}_i +
482: \frac{\tau}{m_0} \sum_{j=1}^{N} \vec{R}_j \\
483: \label{fi1b}
484: \vec{R}'_i &=& \vec{R}_i - k^2\, m_i
485: \,\tau\,\sum_{j=1, j\neq i}^{N}
486: \frac{m_j \left( \vec{r}_i - \vec{r}_j
487: \right)}{\Delta_{ij}^3},
488: \end{eqnarray}
489: where $\vec{r}_i$ stands for $\vec{r}_i(t_0)$,
490: $\vec{r}'_i$ stands for $\vec{r}_i(t_0+\tau)$, and similarly for
491: $\vec{R}_i$, $\vec{R}'_i$.
492:
493: Separating $\cK$ according to Equations \Eqref{partnew}, \Eqref{K0n}, and
494: \Eqref{K1n} offers a possibility of using only two maps for a W-H integrator.
495: {
496: Although our partitioning seems to be more compact and clear
497: than the ``heliocentric-democratic'' scheme, we
498: note that both mappings are practically equivalent as far as the CPU cost
499: is concerned. Yet
500: the rule ``two is a company, three is a crowd''
501: }
502: holds true in the realm of symplectic integrators for perturbed systems:
503: according to the theorem of \citet{Suzuki:91} any symplectic composition method
504: of an order higher than 2 must necessarily involve stages with negative
505: sub-steps that amplify accumulation of roundoff errors. This holds true for a
506: composition of maps derived from splitting the Hamiltonian into any number of
507: terms. However, \citet{LasRob} found a family of methods designed for two-terms
508: perturbed systems with $\cK = \mathcal{A} + \varepsilon \mathcal{B}$ where the
509: negative sub-steps are avoided. Their integrators do not contradict the results
510: of Suzuki: formally they remain second order methods, but in contrast to other
511: W-H methods with local truncation errors $O(\varepsilon \tau^2)$, their errors
512: have a form $O(\varepsilon^2 \tau^2 + \varepsilon \tau^n)$. So, for sufficiently
513: small perturbation $\varepsilon$, the Laskar-Robutel methods may behave like
514: higher order integrators in certain domain of stepsize $\tau$ although no
515: negative substeps were introduced.
516:
517: \subsection{Tangent maps}
518:
519: Whenever a differential correction of initial conditions or the computation of
520: sensitivity indicators is required, the use of tangent maps becomes
521: indispensable. Keplerian map $\Phi_0$ and its associate tangent map can be
522: computed according to a comprehensive recipe by \citet{MikIn:99}. The
523: propagation of a tangent vector
524: \begin{equation}\label{tvdef}
525: \vec{\xi} = \left(%
526: \begin{array}{c}
527: \delta\vec{r} \\
528: \delta\vec{R} \\
529: \end{array}%
530: \right),
531: \end{equation}
532: under the action of the ``kick'' map $\Phi_1$ amounts to multiplying it
533: by the Jacobian matrix $\mathrm{D}\Phi_1$. Resulting expressions are simple:
534: \begin{eqnarray}\label{tv1}
535: \delta\vec{r}' & =& \delta\vec{r} + \frac{\tau}{m_0} \sum_{j=1}^N \delta\vec{R}_j, \\
536: \delta\vec{R}_i' & =& \delta\vec{R}_i + \tau \,k^2\,m_i \sum_{j=1,\,j \neq i}^N
537: \frac{m_j}{\Delta_{ij}^3} \left[ \vec{\delta}_{ij}
538: + \frac{3\,\delta\Delta_{ij}}{\Delta_{ij}}\,(\vec{r}_i-\vec{r}_j) \right],
539: \end{eqnarray}
540: where
541: \begin{equation}\label{dd}
542: \delta\Delta_{ij} = \frac{\left(\vec{r}_j-\vec{r}_i \right) \cdot \vec{\delta}_{ij}}{\Delta_{ij}} ,
543: \qquad
544: \vec{\delta}_{ij} = \delta\vec{r}_j-\delta\vec{r}_i .
545: \end{equation}
546:
547: \subsection{Angular momentum integral}
548:
549: It can be easily demonstrated, that any composition of maps $\Phi_0$ and $\Phi_1$ conserves
550: the angular momentum integral (\ref{GP}). This property is guaranteed by the conservation of $\vec{G}$
551: by each map separately. Recalling that $\Phi_0$ and $\Phi_1$ define \emph{exact} solutions
552: of motion generated by $\cK_0$ and $\cK_1$ respectively, we can simply check that
553: \begin{equation} \label{Gcon}
554: \{\vec{G};\,\cK_0\} = \{\vec{G};\,\cK_1\} = \vec{0}.
555: \end{equation}
556: The proof of Eq.~(\ref{Gcon}) is straightforward. Starting from the definition of
557: $\vec{G}$, we use the linearity of Poisson brackets and the Leibnitz identity to write
558: \begin{eqnarray}
559: % \nonumber to remove numbering (before each equation)
560: \{\vec{G};\,\cK_0\} & = & \sum_{i=1}^N \vec{r}_i \times \{\vec{R}_i;\,\cK_0\}
561: - \vec{R}_i \times \{\vec{r}_i;\,\cK_0\} = \nonumber \\
562: &=& - \sum_{i=1}^N \vec{r}_i \times \frac{\partial\,\cK_0}{\partial \vec{r}_i}
563: + \vec{R}_i \times \frac{\partial\,\cK_0}{\partial \vec{R}_i}.
564: \end{eqnarray}
565: Then we substitute the right-hand sides of Equations (\ref{kep:1}) and (\ref{kep:2}), concluding that
566: all vector products vanish and indeed $ \{\vec{G};\,\cK_0\} = \vec{0}$. A similar procedure demonstrates
567: $ \{\vec{G};\,\cK_1\} = \vec{0}$.
568:
569: Of course, from practical point of view the conservation of $\vec{G}$ is only up to
570: computer roundoff errors.
571:
572: \subsection{Chaoticity indicators}
573: \label{S:MEGNO}
574:
575: To detect unstable motions in the phase space, many numerical tools are
576: available. Concerning the dynamics of close-to-integrable Hamiltonian systems,
577: they can be roughly divided in two classes: spectral algorithms that resolve the
578: fundamental frequencies and/or their diffusion rates
579: \citep{Laskar1993,Nesvorny1997,Michtchenko2001}, and methods based on the
580: divergence rate of initially close phase trajectories, expressed in terms of the
581: Lyapunov exponents \citep{Benettin1980,Froeschle1984}.
582:
583: In this work, among the the spectral tools, we choose the method invented by
584: \citet{Michtchenko2001}; its idea is genuinely simple --- to detect chaotic
585: behavior one counts the number of frequencies in the FFT-spectrum of an
586: appropriately chosen dynamical signal. We deal with conservative Hamiltonian
587: systems; so in a regular case, the spectrum of fundamental frequencies is
588: discrete and we obtain only a few dominant peaks in the FFT spectrum. Chaotic
589: signals do not have well defined frequencies, and their FFT spectrum is very
590: complex. The number of peaks in the spectrum above some noise level $p$
591: (typically, $p$ is set to a few percent of the dominant amplitude) tell us on
592: the character (regular, chaotic) of the of the system.
593:
594: {
595: The method by \citet{Michtchenko2001} does not have as strong theoretical
596: foundations as the Frequency Map Analysis (FMA) by \cite{Laskar1993} or the
597: Fourier Modified Transform (FMT) by \cite{Nesvorny1997} which are considered as
598: rigorous and efficient tools. We did some comparative tests of the later
599: algorithm with MEGNO already \citep{Gozdziewski2004}. Here, we choose the
600: method of \cite{Michtchenko2001} for its appealing simplicity and because we
601: used it in the former papers devoted to the analysis of the RV data. In that
602: way, we can compare the results directly. In our code, the spectral signals
603: analyzed with the FFT are related to canonical Poincar\'e elements, so the
604: fundamental frequencies are well defined. Moreover, we resolve the chaotic and
605: regular signals by comparing the number of significant peaks in the FFT
606: spectrum, thus a very precise determination of the fundamental frequencies is
607: not critical. Actually, in this work we also show that the algorithm has same
608: drawbacks and should be applied with care.
609: }
610:
611: The basic tool to discover exponentially unstable bounded orbits, i.e. chaotic
612: orbits, is the Maximum Lyapunov Characteristic Exponent (MLCE) $\sigma$.
613: Numerical symplectic integration methods are fixed step algorithms, so we can
614: restrict our discussion to the iterations of a discrete map $\vec{\zeta}_n =
615: \Phi^n \vec{\zeta}_0$, that generates a sequence of state vectors
616: $\vec{\zeta}_n$ consisting of coordinates and their conjugate momenta. The
617: direct computation of the MLCE is based on the analysis of the tangent vectors
618: $\vec{\delta}_n$ that evolve under the action of a linear tangent map
619: $\vec{\delta}_n = (D\Phi)^n \vec{\delta}_0$. Asymptotically, the MLCE value is
620: given by
621: \begin{equation} \label{sig}
622: \sigma = \lim_{n \rightarrow \infty} \frac{1}{n} \sum_{k=1}^n
623: \ln \left(\frac{\delta_k}{\delta_{k-1}}\right).
624: \end{equation}
625: If $\sigma$ converges to some positive value, we conclude that the nominal orbit
626: $\vec{\zeta}_n$ and some initially close orbit diverge exponentially at the rate
627: $\exp(\sigma t)$. Two practical difficulties arise when the direct definition
628: (\ref{sig}) is used: the convergence of $\sigma$ is often very slow, and it is
629: difficult to tell how small should be the final value of $\sigma$ to consider it
630: $\sigma=0$.
631:
632: A large variety of methods has been proposed to overcome the problem of slowly
633: convergent MLCE estimates. The authors prefer the so called MEGNO (Mean
634: Exponential Growth factor of Nearby Orbits) indicator proposed by
635: \citet{Cincotta2000} -- that choice is justified by the successful application
636: of this method in our previous works \citep[e.g.,][and the references
637: therein]{BMBW:2005,Gozdziewski2006a}. The definition of MEGNO for a discrete
638: map is \citep{Cincotta2003}
639: \begin{equation}\label{mfm2}
640: Y(n) = \frac{1}{n} \sum_{k=1}^n y(k),
641: \end{equation}
642: where
643: \begin{equation}\label{mfm1}
644: y(n) = \frac{2}{n} \sum_{k=1}^n k\,\ln\left(
645: \frac{\delta_k}{\delta_{k-1}}\right).
646: \end{equation}
647:
648: If the iterates of the discrete map refer to the moments of time
649: separated by the stepsize $h$, the discrete map MEGNO function $Y(n)$ asymptotically
650: tends to
651: \[
652: Y_n = a\,h\,n+b,
653: \]
654: with $a=0, b=2$ for a quasi-periodic orbit, $a=b=0$ for a stable, isochronous
655: periodic orbit, and $a = \frac{1}{2} \sigma,\,b=0$ for a chaotic orbit. Thus we
656: can indirectly estimate the MLCE on a finite time interval, but the weight
657: function $k$ in the sum (\ref{mfm1}) reduces the contribution of the initial
658: part of the tangent vector evolution, when the exponential divergence is to
659: small to be observed behind other linear and nonlinear effects
660: \citep{Morbidelli:book}. Thus, fitting the straight line to the final part of
661: $Y(n)$, we obtain good estimates of $\sigma$ from a relatively shorter piece of
662: trajectory than in the direct MLCE evaluation.
663:
664: In practical application, one can use a more convenient form of
665: Eqs.~(\ref{mfm2}) and (\ref{mfm1}) proposed by \cite{BMBW:2005}
666: \begin{eqnarray} \label{updY}
667: Y(n) & = & \frac{(n-1) \, Y(n-1) + y(n)}{n}, \\
668: \label{updy}
669: y(n) & = &
670: \frac{n-1}{n}\,y(n-1) + 2 \, \ln \left(\frac{\delta_n}{\delta_{n-1}}\right),
671: \end{eqnarray}
672: with the initial setup $y(0)=Y(0)=0$. The fact that only the ratio
673: $\delta_n/\delta_{n-1}$ is significant, as well as the linearity of tangent map,
674: allows to avoid the overflow of $\delta_n$ thanks to occasional normalization of
675: the tangent vector length to $\delta_n=1$ performed after the ratio of
676: $\delta_n/\delta_{n-1}$ has been evaluated.
677:
678:
679: \section{Stability of the HD 37124 planetary system}
680:
681: As a non-trivial application of the presented algorithms and the illustration of
682: difficulties arising in the dynamical analysis of the long-term stability of
683: multiplanet configurations, we choose the HD~37124 extrasolar system. The
684: discovery of two Jovian planets has been announced by \cite{Butler2001} and
685: confirmed by \cite{Vogt2005}. At first, the system seemed to be well modeled by
686: a 2-planet configuration \citep{Butler2001}. However, new observations lead to
687: two-planet fits with $e_c \sim 0.7$ and a catastrophically unstable
688: configuration. Moreover, with the updated RV observations, \citet{Vogt2005}
689: found much better model of 3 planets with similar masses of
690: $\sim0.6$~$m_{\idm{J}}$ in low-eccentric orbits. The best fits have the rms
691: $\sim4$~m/s, in agreement with the internal accuracy of the data. However, the
692: best-fit orbital solution, both in the the kinematic Keplerian model, and in
693: more realistic $N$-body simulation (see Table~1), lies close to the collision
694: line of planets c~and d. Note, that we define the collision line in terms of
695: semi-axes and eccentricities as $a_c (1+e_c) = a_d (1-e_d)$. This line marks
696: the zone in which the mutual interactions of massive companions can quickly
697: destabilize the system.
698:
699: \begin{table}
700: \label{tab:tab1}
701: \caption{
702: The bet-fit astro-centric, osculating Keplerian elements of a stable HD~37124
703: planetary configuration at the epoch of the first observation
704: $t_0$=JD2,451,0420.047. Mass of the parent star is 0.78~$m_{\sun}$. The fit has
705: been refined with GAMP over $\sim 5\cdot 10^4 P_{\idm{d}}$. See
706: \citep{Vogt2005,Gozdziewski2006a} and Fig.~\ref{fig:fig2} for more details.
707: }
708: \centering
709: \begin{tabular}{lccc}
710: \hline
711: Parameter \hspace{1em}
712: & \ \ planet b \ \ & \ \ planet c \ \ & \ \ planet d
713: \\
714: \hline
715: $m \sin i$ [m$_{\idm{J}}$]
716: & 0.62447 & 0.56760 & 0.71194
717: \\
718: $a$ [AU] & 0.51866 & 1.61117 & 3.14451
719: \\
720: $e$ & 0.07932 & 0.15267 & 0.29775
721: \\
722: $\omega$ [deg] & 138.405 & 268.863 & 269.494
723: \\
724: ${\cal M}(t_0)$ [deg]
725: & 259.011 & 109.545 & 124.113
726: \\
727: $\Chi$ & \multicolumn{3}{c}{0.938}
728: \\
729: $V_0$ [m s$^{-1}$] & \multicolumn{3}{c}{7.629}
730: \\
731: rms~[m s$^{-1}$] & \multicolumn{3}{c}{3.39}
732: \\
733: \hline
734: \end{tabular}
735: \end{table}
736:
737: %#[amoeba] chi2-rms-rms1-MEGNO 0.9378 3.3917 4.0763 2.0000
738: %#Better: vr0-chi2-rms-rms1 7.6288128 0.9377849 3.3916522 4.0762586
739: % 0.62447 0.51866 0.07932 90.00000 0.00000 138.40533 259.01067
740: %#
741: % 0.56760 1.61117 0.15267 90.00000 0.00000 268.86257 109.54526
742: %#
743: % 0.71194 3.14451 0.29775 90.00000 0.00000 269.49362 124.11285
744: %#
745: % 0.7800000 7.6288128
746:
747: How to interpret the RV measurements remains an open question. The dynamical
748: long-term stability of the planetary system is the most natural requirement of a
749: configuration consistent with observations. Yet the three-planet model is
750: parameterized by at least 16~parameters, even assuming that the system is
751: coplanar. For that reason the search for the best fits fulfilling the
752: constraints of stability is a difficult task. It can be resolved in different
753: ways. For example, we may try to find dynamically stable solutions in the
754: vicinity of the formal best fit configurations (the latter are often unstable).
755: However, examining the stability of configurations in that neighborhood, we have
756: no reasons to expect that the stable fits are optimal in the statistical sense.
757: Another approach relies on the elimination of unstable fits {\em during} the
758: fitting process, through {\em penalizing} unstable solutions with a large value
759: of $\Chi$. This method, described in \citet{Gozdziewski2006a}, is dubbed GAMP
760: (Genetic Algorithm with MEGNO Penalty). It was shown, that such an approach is
761: particularly useful in modeling resonant or close-to-resonant planetary
762: configurations.
763: {
764: Unfortunately, the algorithm cannot give definite answer when we want to
765: resolve the $\Chi$ shape of strictly stable solutions in detail. The penalty
766: term in the $\Chi$ function relies on a signature of the system stability,
767: expressed through the fast indicator. Due to significant CPU overhead, the fast
768: indicator in the minimizing code can be only calculated over relatively short
769: time, typically $10^3$ orbital periods of the outermost planet. Moreover, the
770: code can converge to unstable best fits that appear stable on that short time
771: scale. Hence, at the end of the search, we have to examine the stability of the
772: individual best fits in the obtained ensemble of solutions, over the time-scale
773: of relevant mean motion and secular resonances.
774: }
775:
776: \begin{figure*}
777: \centerline{
778: \includegraphics[width=5.2in]{fig1.eps}
779: }
780: \caption{
781: Evolution of the HD~37124 system selected best fit related to the theree-body
782: MMR (initial conditions listed in the text). Left: evolution of MEGNO over a
783: short (top) and a long (bottom) time. The straight line is the least-square fit
784: to $Y(t)=(\sigma/2) t +b $. Top right: contact eccentricities during
785: $\sim15$~Myr. Bottom right: the critical argument of the three-body MMR
786: $+1\mbox{b}:-8\mbox{c}:+7\mbox{d}$.
787: }
788: \label{fig:fig1}
789: \end{figure*}
790:
791: \subsection{Long-term stability of the best-fit configurations}
792:
793: The GAMP analysis of the RV data of HD~37124 published by \citet{Vogt2005} was
794: presented in \cite{Gozdziewski2006a}. About 100 of best fit solutions were
795: found yielding $\Chi<1.1$, the rms $\sim4$~m/s, and {\em stable} in the sense
796: that their MEGNO signatures are close to 2 up to $\sim1000-2000$ orbital periods
797: of the outermost planet. Due to heavy CPU requirements, the time-span to
798: resolve MEGNO in the GAMP code cannot be set very long. The short integration
799: time $\sim10^3 P_{\idm{d}}$ allows only to eliminate strongly chaotic
800: configurations, typically leading to collisions between planets and/or with the
801: parent star. The best fit solutions were found in a dynamically active region of
802: the phase space, spanned by a number of low-order mean motion resonances (MMRs)
803: between the two outermost Jovian companions, like 5c:2d, 8c:3d, or 11c:4d (see
804: Fig.~\ref{fig:fig2} and dynamical maps presented in
805: Figs.~\ref{fig:fig3},\ref{fig:fig4},\ref{fig:fig5},\ref{fig:fig6}). In
806: particular, close to the collision lines, the low-order MMRs overlap, giving
807: rise to the global instability zone.
808:
809: Yet we should be aware that two-body MMRs with characteristic time scale $\sim
810: 10^4$--$10^5$ orbital periods of the outermost planet are not the only source of
811: instability in the multi-planet system. Already when we deal with three-planet
812: configurations, the strong instabilities may be generated by three-body MMRs or
813: by long-term secular resonances
814: \citep{Nesvorny1998,Nesvorny1999,Murray1998,Guzzo2006}. In such instance,
815: appropriately longer integration time is necessary to detect the unstable
816: solutions.
817:
818: This issue is illustrated in Fig.~\ref{fig:fig1}. We choose one of the best fits
819: with initial osculating, astrocentric Keplerian elements at the epoch of the first
820: observation in terms of tuples ($m$~[m$_{\idm{J}}$], $a$~[AU], $e$,
821: $\omega$~[deg], ${\cal M}$~[deg]):
822: (0.593, 0.519, 0.0058, 303.360, 95.060),
823: (0.558, 1.615, 0.101, 315.621, 70.279), and
824: (0.690, 3.193, 0.26111, 255.848, 142.886)
825: for planets b, c, and d, respectively. These initial conditions had
826: $\Chi \approx 0.98$ and an rms about $4$~m/s.
827: In the time range covered by the GAMP integrations (and up to $\sim10^4
828: P_{\idm{d}}$, i.e. $\sim60,000$~yr), the configuration appears strictly regular
829: because the indicator quickly converges to~2 (the top-left panel in
830: Fig~\ref{fig:fig1}). Nevertheless, after the transient time, the MEGNO starts to
831: grow linearly at the rate of $\sigma/2 \sim2\cdot 10^{-4} \mbox{yr}^{-1}$ (where
832: $\sigma$ is the MLCE of the solution, see bottom-left panel in
833: Fig.~\ref{fig:fig1}). Actually, after a relatively long time $\sim15$~Myr, the
834: chaotic motion leads to a collision between planets c~and d (the top-right
835: panel) due to a sudden increase of both eccentricities up to 0.6. The
836: elimination of such solutions during an extensive GAMP-like search on a Myrs
837: interval would be very difficult.
838:
839: Looking for the source of such dramatically unstable behavior, we perform the
840: frequency analysis of the orbits with the MFT by \citet{Nesvorny1997}. Denoting
841: the proper mean motions by $n_\mathrm{b},
842: n_\mathrm{c}$, and
843: $n_\mathrm{d}$, respectively, we found that
844: \[
845: n_\mathrm{b} - 8 n_\mathrm{c} + 7 n_\mathrm{d} \approx -0.4^{\circ}/\mbox{yr},
846: \]
847: clearly indicating the three-body MMR of the first order, and we label it with
848: $+1\mbox{b}:-8\mbox{c}:+7\mbox{d}$. The time evolution of the critical argument
849: $\theta=\lambda_\mathrm{b} - 8 \lambda_\mathrm{c} + 7 \lambda_\mathrm{d}$ is
850: illustrated in the bottom-right panel of Fig.~\ref{fig:fig1}. The circulation of
851: the critical angle alternates with libration, indicating the separatrix
852: crossings that explain chaotic evolution.
853:
854: The presented example has inspired us to follow a two-stage procedure in
855: modeling the RV data. First, with a GAMP-like code we look for many best-fit
856: solutions, ideally, approximating the global shape of $\Chi$ and simultaneously
857: stable, at least over a relatively short period of time. At that stage the
858: stability constraints cannot be tight, not only due to significant CPU
859: requirements but also because we should not discard weakly chaotic solutions.
860: Such configurations may be bounded over very long time, longer by orders of
861: magnitude than their Lyapunov time $T_{\idm{L}}=1/\sigma$. In the next step, we
862: either refine the search in a zone bounded by the previously found fits with
863: much longer integration times (still numerically expensive), or we examine each
864: fit with long-term direct integrations and/or evaluate a fast indicator
865: signature, like the MLCE, Spectral Number, or the diffusion of fundamental
866: frequencies.
867:
868: Here, for each solution with $\Chi<1.1$, we computed its MEGNO signature. The
869: integration time span is about of 37~Myr -- long enough to detect the relevant
870: chaotic three-body resonances and strong secular resonances. Here, and in the
871: experiments described later on, we use the SBAB3 integrator scheme by
872: \citet{LasRob}. The time-step is 4~days. The secular periods in the given range
873: of $a_\mathrm{d}$ are quite short, $\sim10^{4}$~yr, nevertheless we can expect that
874: dynamical effects of potentially active secular resonances could be detected
875: after thousands of such characteristic periods, hence counted in
876: $10^6$--$10^7$~yr. Figure~\ref{fig:fig2} illustrates the results. The quality of
877: fits in terms of $\Chi$ is marked by the size of circles (better fits have
878: larger circles). Red (medium grey) circles are for stable, quasi-periodic
879: solutions. In that case the system may be stable over a very long time. Blue
880: (dark grey) circles are for chaotic solutions that led to collisions between
881: planets and that did not survive during the integration time (the integrations
882: are interrupted if any of the eccentricities increases above 0.66). Finally,
883: small yellow (light grey) circles mark all configurations (not necessarily
884: regular) that survived, remaining bounded during the maximal integration time.
885: Clearly, most of solutions with initial $e_\mathrm{d}>0.2$ are both chaotic and
886: unbounded. Nevertheless, some chaotic solutions appear on the borders of stable
887: regions as well. Generally, the distribution of fits gives us a clear image of
888: the border of global instability of the system, relatively far from the
889: collision zone.
890:
891: \begin{figure*}
892: \includegraphics[]{fig2.eps}
893: \caption{
894: The long-term stability of the best-fits obtained in the GAMP search by
895: \citet{Gozdziewski2006a}. The osculating elements at the epoch of the first
896: observation (JD~2,450,420.047) are projected onto the ($a_d,e_d$)-plane (i.e.,
897: the semi-major axis vs the eccentricity of the outermost planet). The quality of
898: fits in terms of $\Chi<1.1$ and rms$<4$~m/s is marked by the size of blue (dark
899: grey) and red (medium grey) circles (larger circles have better fit quality).
900: The best fit, self-consistent Newtonian
901: configuration obtained without stability constraints, in terms of quintuples
902: ($m$~[m$_{\idm{J}}$], $a$~[AU], $e$, $\omega$~[deg], ${\cal M}$~[deg]) at the
903: epoch of the first observation, is
904: (0.619, 0.519, 0.088, 141.91, 257.34),
905: (0.565, 1.663, 0.104, 331.88, 67.71), and
906: (0.732, 2.947, 0.378, 283.33, 95.32)
907: for planets b, c, and d, respectively, with velocity offset $7.53$~m/s. The blue
908: circles are for chaotic solutions that did not survive the integration time of
909: $\sim37$~Myr. The red circles are for regular solutions --- in that case the
910: MEGNO converged to~2. The small yellow (light grey) circles mark configurations
911: that survived the integration. The best {\em stable} fit found is marked with
912: an arrow, its osculating elements are given in Table~1. Some dominant MMRs of
913: planets c and d are marked with dashed vertical lines, according to the third
914: Kepler law, and labeled accordingly. The red curve marks the collision line of
915: the two outermost orbits.
916: }
917: \label{fig:fig2}
918: \end{figure*}
919:
920:
921:
922: \subsection{Fine structure of the phase space}
923:
924: Figure~\ref{fig:fig3} compares the sensitivity of MEGNO and the Spec\-tral
925: Number when we use the same integration time, $\sim 10^5~\mathrm{yr} \approx 1.6
926: \cdot 10^4~P_{\idm{d}}$. In the case of the SN map, we did the FFT on
927: $N=2^{19}$~steps of 64~d, counting the number of spectral lines above $1\%$ of
928: the largest amplitude in the signal of $f(t) = a_{\idm{c}}(t) \exp
929: \mbox{i}\lambda_{\idm{c}}(t)$, where $a_{\idm{c}}$ and $\lambda_{\idm{c}}$
930: denote the {\em contact} semi-major axis and mean longitude of planet~c.
931:
932: Both dynamical maps present the same region of the phase space, in the
933: neighborhood of the best fit. Note, that this particular solution has been
934: refined with GAMP integration over time $5\sim10^4~P_{\idm{d}}$ that is about
935: of 2 orders of magnitude longer than in the set of selected solutions. The
936: resolution of the maps is the same: $480 \times 120$ data points; the map
937: coordinates are usual astrocentric osculating Keplerian elements. Most of
938: best-fits from Fig.~\ref{fig:fig2} lie in the region covered by
939: Fig.~\ref{fig:fig3}.
940:
941: Both maps reveal a number of unstable resonances. Yet the SN map involves some
942: artifacts (\emph{moire}-like patterns) related to a low value of the noise level
943: parameter $p$. Within the same integration time, the MEGNO map reveals
944: relatively more fine details than SN. In particular, the sophisticated border of
945: the collision zone appears to be more sharp and shifted towards smaller
946: $e_{\idm{d}}$. We can also find some fine resonance lines entirely absent in
947: the SN map. For instance, there is a fine structure on the right-hand side of
948: the $8\mbox{c}:3\mbox{d}$~MMR. In order to investigate that instability, we
949: choose the initial condition marked with small crossed circle and labeled
950: with~\textbf{a}. The results of the MFT frequency analysis of this solution tell
951: us that the structure is related to the $+2\mbox{b}:-12\mbox{c}:+3\mbox{c}$ MMR.
952: The time evolution of the related critical argument is illustrated
953: Fig.~\ref{fig:fig8}a.
954:
955: \begin{figure*}
956: \centerline{
957: \includegraphics[width=14.6cm]{fig3.eps}
958: }
959: \caption{
960: Dynamical maps of the Spectral Number (left) and the MEGNO indicator (right)
961: computed in the neighborhood of the best-fit solution to the RV data of HD~37124
962: with the integration time $\sim10^5~\mathrm{yr}$. The elements of the best fit
963: (see Table~1) are labeled with the crossed circle. The rectangle (I) marks the
964: borders of the close-up shown in Fig.~\ref{fig:fig4}. The stability of orbits is
965: color-coded: in both maps, yellow (pale grey) means strongly chaotic and
966: unstable solutions; regular configurations are marked black in the SN-map and
967: dark blue (dark grey) in the MEGNO map.
968: }
969: \label{fig:fig3}
970: \end{figure*}
971:
972:
973: \begin{figure*}
974: \centerline{
975: \includegraphics[width=14.6cm]{fig4.eps}
976: }
977: \caption{
978: Dynamical maps in terms of the Spectral Number (the left panel) and the MEGNO
979: indicator (the right panel) computed in the region marked with small
980: rectangle~I in Fig.~\ref{fig:fig4}. The integration time $\sim3\times 10^5$~yr
981: is equivalent to $\sim4.8 \cdot 10^4$ orbital periods of the outermost planet.
982: The resolution of the maps $500\times120$ data points. The stability of orbits
983: is color-coded, see the caption to the previous figure.
984: }
985: \label{fig:fig4}
986: \end{figure*}
987:
988: Next, we computed close-ups of the dynamical map within the rectangle labeled~I
989: in Fig.~\ref{fig:fig3}. These maps are shown in Fig.~\ref{fig:fig4}. This time
990: we increased $p$ to $5\%$ and the total number of steps has been doubled
991: ($N=2^{20}$) in order to avoid the moire artifacts. But once again the
992: ''concurrent'' MEGNO map (the right panel of Fig.~\ref{fig:fig4}) calculated
993: over the same total time seems to offer a better representation of the phase
994: space. Interestingly, the best fit data (Table~1) seem to lie on the border of a
995: chaotic zone spanned by many overlapping resonances. A close-up of that area,
996: marked with rectangle~II in Fig.~\ref{fig:fig4}, is shown in
997: Fig.~\ref{fig:fig5}. Clearly, even a very small change of parameters of the
998: outermost planet may push the system into a strongly chaotic state. It also
999: illustrates the good performance of the GAMP algorithm that was able to locate
1000: and preserve the fit in an extremely narrow island of stable motions.
1001:
1002: \begin{figure}
1003: \centerline{\includegraphics[width=7cm]{fig5.eps}}
1004: \caption{
1005: Dynamical maps in terms of the MEGNO indicator computed in the region marked by
1006: small rectangle in Fig.~\ref{fig:fig4}. The integration time is $\sim3\times
1007: 10^5$~yr that is equivalent to $\sim4.8 \cdot 10^4$ orbital periods of the
1008: putative outermost planet. The resolution of the plot is $200\times320 $ data
1009: points.
1010: }
1011: \label{fig:fig5}
1012: \end{figure}
1013:
1014: A closer inspection of the area III in Fig.~\ref{fig:fig4} reveals a multitude
1015: of weakly unstable solutions. To show such structures in more detail, we
1016: computed a close-up of that area (Fig.~\ref{fig:fig6}). The resolution of the
1017: maps is $200\times200$ data points, the integration interval is $\sim3\cdot
1018: 10^5~\mathrm{yr} \approx 5\cdot10^4 P_{\idm{d}}$. The time step at the left
1019: column is 16~days and it provides the relative error of the total energy at the
1020: level of $10^{-8}$ . Apparently, in spite of relatively short integration time,
1021: the map uncovers sophisticated structure of the two-body and three-body
1022: resonances. To identify some of them, we choose initial condition marked by
1023: small crossed circles and labeled in the map with \textbf{b}, \textbf{c}, and
1024: \textbf{d}, respectively.
1025:
1026: For initial condition \textbf{b} we found that
1027: $
1028: n_\mathrm{b} - 11 n_\mathrm{c} + 15 n_\mathrm{d} \approx -0.1^{\circ}/\mbox{yr},
1029: $
1030: i.e., indicating three-body MMR $+1\mbox{b}:-11\mbox{c}:+15\mbox{d}$;
1031: for initial condition \textbf{c} we have got
1032: $
1033: 10 n_\mathrm{c} - 27 n_\mathrm{d} \approx -0.3^{\circ}/\mbox{yr},
1034: $
1035: corresponding to the $+10\mbox{c}:-27\mbox{d}$ MMR of the outer
1036: giants; and for initial condition \textbf{d} we have
1037: $
1038: n_\mathrm{b} - n_\mathrm{c} -12 n_\mathrm{d} \approx -0.2^{\circ}/\mbox{yr},
1039: $
1040: indicating the three-body $+1\mbox{b}:-1\mbox{c}:-12\mbox{d}$ MMR, respectively.
1041: All these resonances excite chaotic configurations. That can be demonstrated
1042: through observing the time evolution of their critical angles (see
1043: Fig.\ref{fig:fig8}b,c,d). In all those instances, we found that the libration
1044: alternates with circulation of these angles, hence confirming that the relevant
1045: configurations are close to the resonance separatrices.
1046:
1047: A particularly interesting star-like structure can be be seen around the
1048: initial condition \textbf{d} (the upper-left panel in Fig.~\ref{fig:fig6}).
1049: In that area, the two-body $10\mbox{c}:27\mbox{d}$
1050: MMR and many week three-body resonances are active, for instance,
1051: $ +1\mbox{b}:-11\mbox{c}:+15\mbox{d} \approx -1.35^{\circ}/\mbox{yr}$,
1052: $+1\mbox{b}:-1\mbox{c}:-12\mbox{d} \approx -0.2^{\circ}/\mbox{yr}$,
1053: $+2\mbox{b}:-12\mbox{c}:+3\mbox{d} \approx -1.5^{\circ}/\mbox{yr}$,
1054: $+3\mbox{b}:-13\mbox{c}:-9\mbox{d} \approx -1.7^{\circ}/\mbox{yr}$, and
1055: $+10\mbox{c}:-27\mbox{d} \approx 1.2^{\circ}/\mbox{yr}$.
1056:
1057: One might be tempted to attribute this sophisticated structure to the so called
1058: Arnold web \citep{Cincotta2002}. Indeed, a closer look at the branches of the
1059: web shows new fine details and extremely complex dynamical structure, in that
1060: zone, illustrated in the close-up map around initial condition \textbf{d},
1061: Fig.~\ref{fig:fig7}. But the truth is that this particular structure is mostly
1062: spurious, and it occurred due to an improper choice of the integration step. To
1063: shed more light on that issue, we show the map of the relative error of the
1064: total energy (the left-bottom panel in Fig.~\ref{fig:fig6}). The coincidence of
1065: higher energy error streaks (bottom left) with instability patterns detected by
1066: MEGNO (top left) is not conclusive by itself, but when we recompute both maps
1067: using a smaller time step of 10~days (panels in the right column of
1068: Fig.~\ref{fig:fig6}), we notice that the web patterns of higher MEGNO disappear
1069: (Fig.~\ref{fig:fig6}, top right) and the energy error map significantly flattens
1070: (Fig.~\ref{fig:fig6}, bottom right). We conclude that two additional resonance
1071: lines that crossed at \textbf{d} were generated by the so-called `step-size
1072: resonances' \citep{Rauch1999} between proper frequencies of the system and the
1073: sampling frequency of the constant step integrator. The effect of step-size
1074: resonance in a constant step integrator can be avoided either by using a
1075: sufficiently small integration step or by the application of high-order schemes.
1076: Unfortunately, both approaches lead to more time consuming algorithms.
1077:
1078: As we could expect, the symplectic scheme outperforms the classical integration
1079: algorithms. For instance, we found the the MEGNO code driven by the
1080: Bulirsh-Gragg-Stoer ODEX integrator \citep{Hairer1995} with the relative
1081: accuracy set to $10^{-13}$ requires a similar CPU time, as the Laskar-Robutel
1082: SBAB3 scheme with 4~days step-size, but the former leads to a much larger,
1083: secularly growing energy error (which is larger by 2-3 orders of magnitude).
1084:
1085: \begin{figure*}
1086: \centerline{
1087: \includegraphics[width=14.6cm]{fig6.eps}
1088: }
1089: \caption{
1090: Close-up's of the MEGNO dynamical map shown in Fig.~\ref{fig:fig4} within
1091: rectangle labeled with III, illustrating the fine structure of the phase space.
1092: The integration time $\sim 3\times 10^5$~yr is equivalent to $\sim 4.8 \cdot
1093: 10^4$ orbital periods of the putative outermost planet~d. The resolution is
1094: $200\times 200$ points. Panels in the top are for the MEGNO computed by the
1095: symplectic algorithm: the left panel is for the time-step of 16~days, the right
1096: panel if for the time step of 10~days. The bottom row is for the relatitve error
1097: of the total energy, for the same time steps, respectively.
1098: }
1099: \label{fig:fig6}
1100: \end{figure*}
1101:
1102:
1103: \begin{figure}
1104: \centerline{
1105: \includegraphics[width=7cm]{fig7.eps}
1106: }
1107: \caption{
1108: Close-up's of the MEGNO dynamical map shown in the left panel of
1109: Fig.~\ref{fig:fig6} The integration time $\sim 2\times 10^5$~yr is equivalent
1110: to $\sim 3.2 \cdot 10^4$ orbital periods of the putative outermost planet~d. The
1111: resolution is $200\times 480$ points. The MEGNO is computed by the symplectic
1112: algorithm with the time-step of 16~days.
1113: }
1114: \label{fig:fig7}
1115: \end{figure}
1116:
1117:
1118: \begin{figure*}
1119: \centerline{
1120: \includegraphics[width=5.2in]{fig8}
1121: }
1122: \caption{
1123: Time-evolution of the critical arguments of some resonances illustrated in the
1124: dynamical maps in Figs.~\ref{fig:fig4} and~\ref{fig:fig6}. The panels are
1125: labeled accordingly with initial conditions marked by small crossed circles in
1126: these dynamical maps. See the text for more details.
1127: }
1128: \label{fig:fig8}
1129: \end{figure*}
1130:
1131: \section{Conclusions}
1132:
1133: The use of Poincar\'e variables in the studies of the dynamics of close-to
1134: integrable planetary systems offers many advantages. The variables are canonical
1135: and offer a simple form of a reduced Hamiltonian. The Hamiltonian can be split
1136: into a sum of two separately integrable parts: the Keplerian term and a small
1137: perturbation. As such, it can serve to construct a symplectic integrator based
1138: on any modern composition method, including the recent ones invented by
1139: \citet{LasRob}. The tangent map computed with the same integration scheme
1140: provides an efficient way of computing the estimate of maximal Lyapunov exponent
1141: in terms of relatively recent fast indicator MEGNO. The method proves to be much
1142: more efficient than general purpose integrators (like the Bulirsh-Stoer-Gragg
1143: method). Besides, it provides the conservation of the integrals of energy and
1144: the angular momentum that is crucial for resolving the fine structure of the
1145: phase space. From the practical point of view, the symplectic algorithms are
1146: relatively simple for numerical implementation.
1147:
1148: Using the numerical tools, we investigate the long term stability of extrasolar
1149: planetary system hosted by HD~37124. The orbital parameters in the set of our
1150: best, self-consistent Newtonian fits \citep{Gozdziewski2006a} are in accord
1151: with the discovery paper \citep{Vogt2005}. Nevertheless, the observational
1152: window of the system is still narrow and the derivation of the model consistent
1153: with observations is difficult and, in fact, uncertain. The dynamical maps
1154: reveal that the relevant region of the phase space, in the neighborhood of the
1155: mathematically best fit, is a strongly chaotic and unstable zone. The fitting
1156: algorithm (GAMP) that relies on eliminating strongly unstable fits founds
1157: solutions with a similar quality [in terms of $\Chi$] that yields the formal
1158: solution. Moreover, they are shifted towards larger semi-major axes and much
1159: smaller eccentricities of the outermost planet. The orbital evolution of two
1160: outer planets is confined to a zone spanned by a number of low-order two-body
1161: and three-body MMRs. In particular, the three-body MMRs may induce very unstable
1162: behaviors that manifest themselves after many Myrs of an apparently stable and
1163: bounded evolution. To deal with such a problem, the stability of the best fits
1164: should be examined over a time-scale that is much longer than the one required
1165: when only the two-body MMRs are considered. In accord with the dynamical maps,
1166: the stable fits to the RV of HD~37124 should have small eccentricity of the
1167: outermost planet~d, not larger than 0.2-0.3. Moreover, the stable
1168: configurations of the HD~37124 system are puzzling. The best-fit mathematical
1169: three-planet model is surprisingly distant, in the phase space of initial
1170: conditions, from the zone of stable solutions consistent with the RV. It remains
1171: possible that other bodies are present in the system and the three-planet model
1172: is not adequate to explain the RV variability, in spite that it provides
1173: apparently perfect fits. Yet, isolated initial conditions or even sets of
1174: best-fit solutions do not provide a complete answer on the system configuration.
1175: Then the fast indicator approach is essential and helpful to resolve the
1176: dynamical structure of the phase space.
1177:
1178: The results of our experiments confirm and warn that all numerical methods should be
1179: applied with great care. All symplectic methods are constant step integrators.
1180: In that case one should be cautious about the possibility of generating spurious
1181: resonance webs. A proper way to avoid them is to repeat computations with a
1182: different integration step in order to detect step-dependent patterns.
1183:
1184: \section{Acknowledgments}
1185: We thank David Nesvorn{\'y} for a review and comments that improved that
1186: manuscript. This work is supported by KBN Grant No. 1P03D-021-29.
1187:
1188: \bibliographystyle{mn2e}
1189: \bibliography{biblio}
1190: \end{document}
1191: