0803.3246/ms.tex
1: \documentclass[useAMS,usenatbib,usegraphicx]{mn2e}
2: %_______________________________________________________________________________
3: \usepackage{pslatex}
4: \usepackage{amsmath,amsfonts}
5: \newcommand{\Hfull}{\mathcal{H}}
6: \newcommand{\Hsec}{\mathcal{H}_{\idm{sec}}}
7: \newcommand{\HsecJ}{\mathcal{H}^J_{\idm{sec}}}
8: \newcommand{\Hpert}{\mathcal{H}_{\idm{pert}}}
9: \newcommand{\HpertJ}{\mathcal{H}^J_{\idm{pert}}}
10: \newcommand{\HKepl}{\mathcal{H}_{\idm{kepl}}}
11: \newcommand{\Hint}{\mathcal{H}_{\idm{int}}}
12: \newcommand{\Rsec}{\mathcal{R}_{\idm{sec}}}
13: \newcommand{\R}{\mathcal{R}}
14: \def\vec#1{{\mathbf{#1}}}
15: \def\idm#1{{\mbox{\scriptsize #1}}}
16: \def\rev#1{{#1}}
17: \voffset=-0.8cm
18: %_______________________________________________________________________________
19: \title[Secular planetary problem]
20: {A secular theory of coplanar, non-resonant planetary system}
21: \author[C. Migaszewski and K. Go\'zdziewski]{Cezary Migaszewski$^{1}$\thanks{E-mail:
22: c.migaszewski@astri.uni.torun.pl} and Krzysztof Go\'zdziewski$^{1}$\footnotemark[1]\thanks{E-mail:
23: k.gozdziewski@astri.uni.torun.pl}\\
24: $^{1}$Toru\'n Centre for Astronomy, Nicolaus Copernicus University, 
25: Gagarin Str. 11, 87-100 Toru\'n, Poland}
26: \begin{document}
27: %_______________________________________________________________________________
28: \date{Accepted 2008 May 7.  Received 2008 May 2; in original form 2008 February 1}
29: \pagerange{\pageref{firstpage}--\pageref{lastpage}} \pubyear{2008}
30: \maketitle
31: \label{firstpage}
32: %_______________________________________________________________________________
33: \begin{abstract}
34: %_______________________________________________________________________________
35: We present the secular theory of coplanar $N$-planet system, in the absence of
36: mean motion resonances between the planets. This theory relies on the averaging 
37: of a perturbation to the two-body problem over the mean longitudes.  We expand
38: the perturbing Hamiltonian in Taylor series with respect to the ratios of
39: semi-major axes  which are considered as small parameters, without  direct
40: restrictions on the eccentricities. Next, we average out the resulting series
41: term by term. This is possible thanks to a particular but in fact quite
42: elementary choice of the integration variables. It makes it possible to avoid
43: Fourier expansions of the perturbing Hamiltonian. We derive high order
44: expansions of the averaged secular Hamiltonian (here, up to the order of 24)
45: with respect to the semi-major axes ratio. The resulting secular theory is a
46: generalization of the octupole theory.  The analytical results are compared with
47: the results of numerical (i.e., practically exact) averaging. We estimate
48: the convergence radius of the derived expansions, and we propose a further
49: improvement of the algorithm. As a particular application of the method, we 
50: consider the  secular dynamics of three-planet coplanar system. We focus on
51: stationary solutions in the HD~37124 planetary system.
52: \end{abstract}
53: 
54: %_______________________________________________________________________________
55: \begin{keywords}
56: %_______________________________________________________________________________
57: celestial mechanics -- secular dynamics -- analytical methods -- stationary solutions -- 
58: extrasolar planetary systems -- stars:~HD~37124
59: \end{keywords}
60: 
61: %_______________________________________________________________________________
62: \section{Introduction}
63: %_______________________________________________________________________________
64: The recent discoveries of extrasolar planetary systems bring new and interesting
65: problems regarding their dynamical stability and long-term evolution. At
66: present, at least 30 multi-planet systems are known and their number is still
67: growing, thanks to refined techniques of observations. Surprisingly, the orbital
68: parameters of these systems are very different from those typical in the Solar
69: System  architecture --- large planetary masses and  eccentricities are common.
70: Simultaneously, these systems usually are compact. Likely, this property is a
71: consequence of  the observational selection. The most effective detection
72: techniques, like the radial velocity observations, rely on indirect effects of
73: mutual interactions between  planets and their host star. Many of the known
74: multi-planet systems are supposed to be involved in short-term mean motion
75: resonances (MMRs). However, there are also configurations with relatively well
76: separated orbits. In that case the secular interactions may lead to interesting
77: dynamical phenomena.
78: 
79: \rev{
80: To study the long term dynamics of planetary systems, different analytical and
81: numerical techniques are used. The analytical approach is much more effective in
82: the investigations of global, qualitative dynamics than widely applied numerical
83: techniques (including fast indicators or direct numerical integration of the
84: equations of motion). The numerical experiments provide only limited (or local)
85: information on the dynamical features of the studied configurations. Usually,
86: the interpretation of the results of massive calculations can be problematic
87: without solid theoretical background. In contrast, analytical techniques offer
88: much deeper insight into qualitative properties of motion. The analytical
89: approach makes it possible to explore large volume of the phase space. This is
90: crucial for the dynamical studies of extrasolar planetary systems detected
91: during short time of observations. These observations have relatively large
92: errors, they are typically irregularly sampled or degenerated (in the sense that
93: they can provide only limited information on the system state, like the radial
94: velocity technique). This leads to poorly determined or unconstrained orbital
95: and physical parameters of the detected systems. In that case, the analytical
96: theories help us to investigate and/or to detect global properties of the
97: solutions to the equations of motion in  observationally permitted ranges of the
98: parameters. We can investigate in detail certain families of these solutions,
99: their bifurcations and stability. Examples of such solutions are stationary
100: solutions (equilibria) or periodic orbits that build a skeleton of the phase
101: space. Investigating these families, we follow the classic methodology invented by
102: Poincar\'e.  The analytical theories are milestones for detailed
103: numerical studies of particular aspects of the dynamics. Hence, their constant
104: development is always desirable.
105: }
106: 
107: Moreover, due to extreme parameters of the studied configurations, the classic
108: planetary theory developed so far is often too week. For instance, the classic
109: Lagrange-Laplace theory \citep{Murray2000} designed as a model of the secular
110: dynamics of planets in the Solar System fails in the case of large
111: eccentricities and inclinations. Hence, new analytical  and semi-analytical
112: theories are recently developed, breaking the limitations of the classic
113: approach.  One of the most effective techniques for studying the secular 
114: dynamics of extrasolar systems  has been recently invented by
115: \cite{Michtchenko2004} and further developed in \citep{Michtchenko2006}. These
116: papers are devoted to a study of two-planet configurations. In this work, we
117: consider an analytical secular theory of a coplanar system of $N$-planets (point
118: masses) under assumption that the  orbital configurations are not involved in
119: strong mean-motion resonances and that they are far from collisions zones of
120: orbits. We calculated the averaged perturbation in the form of  power series
121: with respect to the semi-major axes ratios up to very high order (equal to 24 in
122: the present work). These expansions have no  explicit limits on the
123: eccentricities provided that the non-resonance condition is satisfied. Our
124: development is elementary and is based on very basic properties of the Keplerian
125: motion.  Although  it concerns the two-planet system, we show that it can be
126: easily generalized for the case of $N$-planet configurations. Hence, the theory
127: can be regarded  not only as an attempt to improve the secular theories for
128: two-planet systems in
129: \citep[e.g.,][]{Gallardo2005,Henrard2005,Libert2005,Libert2006,Ji2007,Veras2007},
130: relying on the classic expansion of the perturbing Hamiltonian in eccentricities
131: \citep{Murray2000,Ellis2000} or the octupole theory that makes use on the
132: averaging of the low-order expansion of the perturbing function in the
133: semi-major axes ratio \citep{Ford2000,Blaes2002,Lee2003}.  We try to reduce the
134: limitations of the classic theory of non-resonant systems. 
135: 
136: The plan of this paper is as follows. In Section~2, we describe a general model
137: of a coplanar configuration of $N$ planets. We introduce  the expansion of
138: perturbing Hamiltonian and a very simple and basic algorithm of its averaging.
139: We compare the results of the method with the  outcome of the octupole theory,
140: and we discuss some subtle differences between these theories. We also present
141: the results of the tests of the expansion, taking as examples a few known
142: multi-planet configurations which apparently fit well in the framework of the
143: non-resonant secular theory. The exact semi-numerical method is helpful to
144: determine absolute bounds of the validity of the analytic approach. We also
145: outline a further improvement of the averaging algorithm.
146: In Section~3 we
147: construct the secular model of the three-planet system and we perform a
148: preliminary study of the secular dynamics of a few known three-planet
149: configurations. In particular, we focus on the equilibria in the secular
150: problem,
151: and we found interesting stationary solutions in the extrasolar system
152: HD~37124~\citep{Vogt2005,Gozdziewski2007}. 
153: 
154: %_______________________________________________________________________________
155: \section{The secular dynamics of a multi-planet system}
156: %_______________________________________________________________________________
157: The Hamiltonian of a multi-planet system with respect to canonical Poincar\'e 
158: variables \citep[see, e.g.,][]{Laskar1995,Michtchenko2004}  can be expressed by
159: a sum of two terms,
160: \begin{equation}
161: \Hfull = \HKepl + \Hpert,
162: \label{Hfull}
163: \end{equation}
164: where
165: \begin{equation}
166: \HKepl = \sum_{i=1}^{N} {\bigg( \frac{\mathbf{p}_i^2}{2 \beta_i} 
167: - \frac{\mu_i \beta_i}{r_i} \bigg)}
168: \label{HKepl}
169: \end{equation}
170: is for the integrable part comprising of the direct sum of the relative,
171: Keplerian motions of $N$ planets and the host star. Here, the dominant point
172: mass of the star is ${m_0}$, and $m_i \ll m_0$, $i=1,\ldots,N$ are the point
173: masses of the $N$-planets.  For each planet--star pair we define the mass
174: parameter ${\mu_i=k^2~(m_0+m_i)}$ where $k$ is the Gauss gravitational constant,
175: and ${\beta_i=(1/m_i+1/m_0)^{-1}}$ are the so called reduced masses. Due to
176: mutual interactions between the planets, the Keplerian part is perturbed by a
177: function $\Hpert$, 
178: \begin{equation}
179: \Hpert \equiv \R = \sum_{i=1}^{N-1} \sum_{j>i}^{N} {\bigg(
180:  - \underbrace{\frac{k^2 m_i m_j}{\Delta_{i,j}}}_{\textrm{\small direct part}} +
181: \underbrace{\frac{\mathbf{p}_i \cdot \mathbf{p}_j}{m_0}}_{\textrm{\small indirect
182: part}}\bigg)},
183: \label{Hpert}
184: \end{equation}
185: where  ${\mathbf{r}_i}$ are for the position vectors of the planets relative to
186: the star,  ${\mathbf{p}_i}$ are for their conjugate momenta relative to the {\em
187: barycenter} of the whole $(N+1)$-body system, 
188: ${\Delta_{i,j}=\|\mathbf{r}_i-\mathbf{r}_j\|}$ denote the relative distance
189: between planets $i$  and $j$. It is well known that even in the simplest case of
190: three  point masses (the star and two mutually  interacting planets), the
191: problem is non-integrable and is not possible to obtain its exact analytical
192: solutions. In practice, such solutions can be only derived in the form of
193: approximations  derived by means of different perturbation techniques
194: \citep[e.g.,][]{Murray2000,Morbidelli2003,FerrazMello2007}. 
195: 
196: To apply the canonical perturbation theory, we first transform $\Hfull$ to the
197: following form:
198: \begin{equation}
199:  \Hfull(\vec{I},\vec{\phi}) = \HKepl(\vec{I}) + 
200:  \Hpert(\vec{I},\vec{\phi}),
201: \label{eq:poincare} 
202: \end{equation}
203: where $(\vec{I},\vec{\phi})$ stand for the action-angle variables, and
204: $\Hpert(\vec{I},\vec{\phi}) \sim \epsilon \HKepl(\vec{I})$,  where $\epsilon \ll
205: 1$ is a small parameter. In the absence of this perturbation, the system is 
206: trivially integrable. However, with the perturbation added, the dynamics of the 
207: full system become extremely complex.  In the realm of the Hamiltonian canonical
208: theory, the approximate, analytical solutions to this problem may be derived by
209: an expansion of the perturbation with respect to the small parameter and by
210: subsequent simplification of the lowest order terms by means of appropriate
211: canonical contact transformations. This idea of Delaunay  appears in many
212: ``incarnations''. One of its first novel realizations is known as the von~Zeipel
213: method \citep[e.g.,][]{Brumberg1995}.  Much more improved version of this
214: technique that not require series inversion  has been invented by
215: \cite{Hori1966} and next refined by \cite{Deprit1969}.  This theory is well
216: known in the literature as the Lie-Hori-Deprit method. For an excellent review
217: of these methods see a monograph by \cite{FerrazMello2007}. Another method of
218: seeking for approximate solutions to Eq.~\ref{eq:poincare} relies directly on
219: the averaging proposition \citep[see, e.g.,][]{Arnold1993}. Usually, the
220: canonical angles $\vec{\phi}$ can be divided onto two classes: {\em fast} and
221: {\em slow} ones. By averaging the perturbing part with respect to the fast
222: angles over their periods, we obtain the secular perturbation Hamiltonian which
223: does not depend on these fast angles. Simultaneously,  their conjugate momenta
224: become integrals of the secular problem. In the planetary model with a dominant
225: stellar mass, we have two natural time-scales of motion: the orbital motion of
226: the planets and a slow evolution of their orbits. Assuming that no strong  mean
227: motion resonances are present, and the system is far enough from collisions, 
228: the averaging makes it possible to reduce the number of the degrees of freedom
229: and to obtain qualitative information on the long-term changes of the slowly
230: varying orbital elements (i.e., on the slow angles and their conjugate momenta).
231: 
232: To apply each one of these methods,  the Hamiltonian of the $N$-planet system
233: should be first transformed to the required form, Eq.~\ref{eq:poincare}. This
234: can be accomplished by expressing it with  respect to the canonical Poincar\'e
235: elements \citep{Murray2000,Michtchenko2004}:
236: \begin{eqnarray}
237: {l_i \equiv \lambda_i}, & \quad {L_i=\beta_i~\sqrt{\mu_i~a_i}},\nonumber\\
238: {g_i \equiv -\varpi_i}, & \quad  {G_i=L_i~(1 - \sqrt{1-e_i^2})},\\
239: {h_i \equiv -\Omega_i}, & \quad{H_i=L_i \sqrt{1-e_i^2}~(1 - \cos~I_i)},\nonumber
240: \label{delaunay}
241: \end{eqnarray}
242: where $\lambda_i$ are the mean longitudes, $a_i$ stand for canonical semi-major
243: axes,  $e_i$ are for the eccentricities,  $I_i$ denote inclinations, 
244: \rev{$\varpi_i$ are the longitudes of pericenter},  and $\Omega_i$ denote the
245: longitudes of the  ascending node. We note that 
246: \rev{
247: the transformation between the canonical orbital elements of Poincar\'e, $a_i$,
248: $e_i$, $I_i$, $\varpi_i$, $\Omega_i$  and associated  Cartesian coordinates and
249: momenta may be derived by the formal two-body transformation between classic
250: (astro-centric) Keplerian elements and the Cartesian coordinates
251: \citep[e.g.,][]{Morbidelli2003,Ferrazmello2006}.
252: }
253: Moreover, in the settings adopted here, the rectangular coordinates and momenta
254: are understood  through the Cartesian positions of planets relative to the star,
255: and, according to the definition of the Poincar\'e variables, the canonical
256: momenta are taken relative to the {\em barycenter} of the system. 
257: 
258: The $N$-body Hamiltonian expressed in terms of the Poincar\'e variables has the
259: form of:
260: \[
261: \Hfull = -\sum_{i=1}^N \frac{\mu_i^2 \beta^3_i}{2 L_i^2} +
262: \Hpert\underbrace{(L_i,l_i,G_i,g_i,H_i,h_i)}_{i=1,\ldots,N}.
263: \]
264: In this Hamiltonian, $l_i$ play the role of the fast angles. In the absence of
265: strong MMRs,  these angles can be eliminated by the following
266: averaging formulae:
267: \begin{equation}
268: \Hsec=\frac{1}{(2 \pi)^N}\underbrace{\int_0^{2 \pi} \ldots 
269: \int_0^{2\pi}}_{i=1,\ldots,N}{\Hpert \, d\lambda_1 \ldots d\lambda_N}.
270: \label{eq:secular}
271: \end{equation}
272: (As we see below, it can be also applied for some selected pairs of planets).
273: Hence, the conjugate momenta $L_i$ become integrals of the secular model and the
274: Keplerian part is a constant that does not contribute to the equations of
275: motion. However, the calculation of the multiple integral is quite a difficult
276: task which is a central part of the problem.  We try to solve it with quite
277: basic mathematical properties of the Keplerian osculating orbits.
278: 
279: \rev{
280: We should keep in mind that the averaging of the secular Hamiltonian in the
281: problem over {\em Keplerian} motions implies truncation of the perturbation
282: to first order in the masses (more generally, to the first order in the
283: perturbation parameter $\epsilon$). For large mutually interacting planets
284: (or binary stars), the deviations of the true orbits from the Keplerian
285: approximation during the orbital period may become significant, and the
286: secular theory may fail. Nevertheless, it is a common drawback of the idea
287: behind Eq.~{\ref{eq:secular}}. For the same reason, the classic perturbation
288: techniques usually fail in the case of close encounters between the planets. 
289: In such an instance, some other criteria helping to explore the stability
290: regions may be applied, for example, the Hill stability criterion
291: \citep{Marchal1982,Gladman1993,Barnes2006,Michtchenko2008,Barnes2008}.
292: }
293: 
294: %_______________________________________________________________________________
295: \subsection{The indirect part of the disturbing function}
296: %_______________________________________________________________________________
297: We start with the averaging of the {\em indirect} part of the disturbing
298: Hamiltonian.
299: \rev{
300: The result of this averaging  can be found in \cite{Brouwer1961}, nevertheless,
301: to make this paper self-consistent, we present
302: the calculations in detail. 
303: }
304: The indirect part
305: is a scalar product of canonical momenta $\mathbf{p}_i$, which
306: have the form of:
307: \begin{equation}
308: \mathbf{p}_i = {\beta'}_i \mathbf{\dot{r}}_i - \sum_{j \neq i} {\frac{m_i m_j}{M}
309: \mathbf{\dot{r}}_j},
310: \end{equation}
311: where ${\beta'}_i = \left[1/m_i + 1/(M-m_i)\right]^{-1}$ and $M$ is the  total
312: mass of the system. The scalar product $\mathbf{p}_i \cdot \mathbf{p}_j$ 
313: includes terms of the type of $\mathbf{\dot{r}}_i \cdot \mathbf{\dot{r}}_j$.
314: Moreover,  each product $\mathbf{p}_i \cdot \mathbf{p}_j$ depends on all 
315: astro-centric velocities of the planets,  $\mathbf{\dot{r}}_i$ ($i=1,\ldots,N$).
316: Apparently, to average out the indirect part of the disturbing function, we must
317: compute multiple integral over all mean longitudes $\lambda_i$ ($i=1,\ldots,N$).
318: In fact, this integral can be reduced to a sum of double integrals computed for
319: all pairs of planets, i.e., we average out expressions of the form of 
320: $\mathbf{\dot{r}}_i \cdot \mathbf{\dot{r}}_j$. The result is the following:
321: \begin{equation}
322: \frac{1}{(2\pi)^2} \int_{0}^{2 \pi} \int_{0}^{2 \pi} {\mathbf{\dot{r}}_i
323: \cdot \mathbf{\dot{r}}_j ~d\mathcal{M}_i d\mathcal{M}_j} = 
324: \delta_{i,j}~a_i^2~n_i^2,
325: \end{equation}
326: where $n_i$ denote mean motions of the planets and $\delta_{i,j}$   stands for
327: the Kronecker delta.   Note, that the averaging over the mean anomalies gives
328: the same results as the averaging over  the mean  longitudes under the condition
329: that the integration limits are set to  $0$ and $2\pi$, respectively. Clearly,
330: the indirect part of the disturbing function does not contribute to  the secular
331: dynamics of the system  because it depends on $L_i$ only \citep{Brouwer1961},
332: \cite[see also][]{Michtchenko2004}. We note that this result is exact as far as
333: the assumptions of the averaging principle are fulfilled (we are far enough from
334: the MMRs and there are present  two different time-scales in
335: the problem).
336: 
337: %_______________________________________________________________________________
338: \subsection{The direct part of the disturbing function}
339: %_______________________________________________________________________________
340: Now we have a more difficult problem to resolve. For the $N$-planet system, the
341: averaged  direct part of the disturbing function has the form of:
342: \begin{equation}
343: \Hsec =
344: \sum_{i=1}^{N-1}  \sum_{j>i}^{N}{\Hsec^{(i,j)}},
345: \label{eq:secular2}
346: \end{equation}
347: where the multiple integral Eq.~(\ref{eq:secular})  over all mean anomalies is
348: reduced to a sum of secular Hamiltonians describing mutual interactions between
349: all pairs of planets; i.e., for each pair $(i,j)$,  where $i<j$ and $a_i<a_j$,
350: we have:
351: \begin{equation}
352: \Hsec^{(i,j)} = 
353:  \frac{1}{(2\pi)^2} \int_0^{2\pi} \int_0^{2\pi} 
354: -{\frac{k^2 m_i m_j}{\Delta_{i,j}} d\mathcal{M}_i d\mathcal{M}_j}.
355: \label{eq:rsec}
356: \end{equation}
357: Hence, the secular model of $N$-planet system can be reduced to a simple sum of
358: two-planet Hamiltonians.  
359: 
360: Now, we compute the double integral for a fixed pair of planets  $i$ and $j$.
361: The distance between these planets is the following:
362: \begin{equation}
363: \Delta_{i,j} = \sqrt{r_i^2 + r_j^2 - 2 r_i r_j \cos{\psi_{i,j}}},
364: \label{squareroot1}
365: \end{equation}
366: where $\psi_{i,j}$ is the angle between vectors $\mathbf{r}_i$ and
367: $\mathbf{r}_j$:
368: \begin{equation}
369: \cos{\psi_{i,j}} = \frac{\mathbf{r}_i \cdot \mathbf{r}_j}{r_i r_j} = 
370: \frac{x_i x_j + y_i y_j}{r_i r_j}.
371: \end{equation}
372: This formulae can be rewritten to:
373: \begin{equation}
374: \Delta_{i,j} = r_j \sqrt{1 - 2 \frac{1}{r_j} \Big(x_i \frac{x_j}{r_j} 
375: + y_i\frac{y_j}{r_j}\Big) + 
376: \Big(\frac{r_i}{r_j}\Big)^2}.
377: \label{eq:squareroot2}
378: \end{equation}
379: According to the Kepler problem theory, $r_i$ and $r_j$ may be expressed through
380: the eccentric anomaly, $E$, or, equivalently, through the true anomaly, $f$. We
381: write down appropriate expressions  for planet $i$ and $j$, respectively: 
382: \[
383: r_i = a_i (1 - e_i \cos{E_i}), \quad
384: r_j = \frac{a_j (1 - e_j^2)}{1 + e_j \cos{f_j}}.
385: \]
386: Here, $E_i$ is the eccentric anomaly of the inner planet in the
387: selected pair of interacting bodies,  and $f_j$ is the true anomaly 
388: of the outer planet. 
389: Moreover:
390: \[
391: \frac{x_j}{r_j} = \cos\left({f_j + \Delta{\varpi_{i,j}}}\right), 
392: \quad \frac{y_j}{r_j} = \sin\left({f_j + \Delta{\varpi_{i,j}}}\right),
393: \]
394: where $\Delta\varpi_{i,j} \equiv \left(\varpi_j - \varpi_i\right)$,
395: and
396: \[
397: x_i = a_i (\cos{E}_i - e_i), \quad y_i = a_i \sqrt{1 - e_i^2} \sin{E_i}.
398: \]
399: The dependence of the {\em two-body formulae} on $\Delta\varpi_{i,j}$ may seem
400: strange. However, when we investigate the co-planar system of two  particular
401: planets, we are free to choose the reference frame
402: because the mutual interaction between these planets depend only on their  {\em
403: relative} orbital phases.  To be more specific, we must calculate the distance
404: between planets $i$ and $j$, or
405: the scalar product $\vec{r}_i \cdot \vec{r}_j \equiv r_i r_j
406: \cos{\psi_{i,j}}$ (Eq. \ref{squareroot1}). That is obviously independent on the
407: reference frame. In general, we could write: 
408: \begin{equation} 
409: \vec{r}_i \cdot \vec{r}_j  =
410: {\mathbb A}_i \vec{r}_i\big|_{{\cal F}_i} 
411: \cdot {\mathbb A}_j \vec{r}_j\big|_{{\cal F}_j}, 
412: \end{equation} 
413: where ${\cal F}_{i,j}$ are the orbital reference frames  of the inner and outer
414: planet, respectively,  matrices ${\mathbb A}_i, {\mathbb A}_j$ represent
415: Eulerian rotations of ${\cal F}_{i,j}$ to a common reference frame (${\cal
416: F}$) for both orbits. Because this  frame may be chosen freely,  we fix the
417: $x$-direction of the common frame along the apsidal line of the inner planet
418: (still, for a particular pair of planets).  Then ${\mathbb A}_i \equiv
419: {\mathbb E}$ and ${{\cal F}_j}$ must be rotated by angle $\Delta{\varpi_{i,j}}$.
420: That can be repeated for each pair of planets in multi-planet system because the
421: secular Hamiltonian is represented by a sum of formally independent two-planet
422: terms.
423: 
424: 
425: Finally, the inverted distance between planets $i,j$ can be expressed
426: as follows:
427: \[
428: \frac{1}{\Delta_{i,j}} = \frac{1 + e_j \cos{f_j}}{a_j (1 - e_j^2)} 
429: \Big[A \alpha_{i,j}^2 -2 B \alpha_{i,j} +1\Big]^{-1/2},
430: \]
431: where $ \alpha_{i,j} \equiv {a_i}/{a_j} < 1$, and
432: \begin{eqnarray}
433: A &\equiv& \frac{(1 + e_j \cos{f_j})^2 (1 - e_i \cos{E_i})^2}{(1 - e_j^2)^2}, \\
434: B &\equiv&  \frac{1 + e_j \cos{f_j}}{1 - e_j^2} \Big[(\cos{E_i} - e_i) 
435: \cos\left({f_j + \Delta{\varpi_{i,j}}}\right) +\\
436: &+& \sqrt{1-e_i^2} \sin{E_i} \sin\left({f_j + \Delta{\varpi_{i,j}}}\right)
437: \Big] \nonumber.
438: \end{eqnarray}
439: Now we underline that the position of the inner planet is  given through the
440: {\em eccentric} anomaly while the position of the outer planet  is given with
441: respect to the {\em true} anomaly. The formulae under the square root are
442: expressed by a polynomial of trigonometric functions and, as we can see below,
443: that is critically important property making it possible to calculate  the
444: integral in Eq.~\ref{eq:rsec}.
445: 
446: Now, we expand the inverse of the distance between planets $i$ and $j$ in  
447: Taylor series with respect to small parameter $\alpha_{i,j}$. The series are
448: evaluated around $\alpha_{i,j}=0$ as follows:
449: \begin{equation}
450: \frac{1}{\Delta_{i,j}} = \frac{1 + e_j \cos{f_j}}{a_j (1 - e_j^2)} \sum_{l=0}^{\infty}
451: \bigg[\frac{1}{l!} \frac{d^l \mathcal{D}}{d\alpha_{i,j}^l} \bigg|_{\alpha_{i,j}=0} 
452: \alpha_{i,j}^l \bigg],
453: \label{eq:oneoverD}
454: \end{equation}
455: where
456: \begin{equation}
457: \mathcal{D} = \Big[A \alpha_{i,j}^2 -2 B \alpha_{i,j} +1\Big]^{-1/2}.
458: \label{eq:eqf}
459: \end{equation}
460: As the final result of this  expansion, we obtain a polynomial with respect to 
461: trigonometric functions of the anomalies which has the general form of:
462: \begin{equation}
463: \frac{1}{\Delta_{i,j}} = \sum_{\mathbf{p}} \Big[C_{\mathbf{p}} (\cos{E_i})^{p_1} 
464: (\sin{E_i})^{p_2} 
465: (\cos{f_j})^{p_3} (\sin{f_j})^{p_4}\Big].
466: \label{eq:oneoverDgeneral}
467: \end{equation}
468: Here, $\vec{p} \equiv (p_1,p_2,p_3,p_4) \in {\mathbb N}^4$  is a vector of
469: natural numbers and  $C_{\vec{p}}$  are coefficients depending on eccentricities
470: and semi-major axes. One more step is still necessary. The integral in 
471: Eq.~\ref{eq:rsec} must be computed with respect to the {\em mean} anomalies. 
472: Here, the classic theories make use on $\sin f$ and $\cos f$ expressed through
473: Fourier series of  the mean anomalies with coefficients dependent on the
474: eccentricities. 
475: 
476: However, we found that it is possible to avoid these expansions. Again, using the basic formulae
477: of the Keplerian motion, we perform a formal change of variables in Eq.~\ref{eq:rsec} with:
478: $d\,\mathcal{M}_i = \mathcal{I}_i dE_i$ and $d\,\mathcal{M}_j = \mathcal{J}_j df_j$, where
479: functions $\mathcal{I}_i \equiv  \mathcal{I}_i(E_i,e_i)$  and $\mathcal{J}_j \equiv
480: \mathcal{J}_j(f_j,e_j)$ are defined with:
481: \begin{equation}
482: \mathcal{I}_i(E_i,e_i) = 1 - e_i \cos{E_i}, \quad
483: \mathcal{J}_j(f_j,e_j) = \frac{\left(1-e_j^2\right)^{3/2}}{\left(1 + e_j \cos{f_j}\right)^2}.
484: \label{eq:ij}
485: \end{equation}
486: After this change of variables, the average in Eq.~\ref{eq:rsec} is equivalent to
487: calculation of the following double integral:
488: \begin{equation}
489: \Hsec^{(i,j)} = 
490:  \frac{1}{(2\pi)^2} \int_0^{2\pi} \int_0^{2\pi} 
491:                   -k^2 m_i m_j \frac{1}{\Delta_{i,j}} \mathcal{I}_i(E_i,e_i) 
492:    \mathcal{J}_j(f_j,e_j) \, d E_i \, d f_j,
493: \label{eq:rsecl}
494: \end{equation}
495: where 
496: ${\Delta^{-1}_{i,j}} \equiv \Delta^{-1}_{i,j}(a_i,a_j,e_i,e_j,E_i,f_j)$.
497: 
498: Fortunately, functions $\mathcal{I}_i$ are again  trigonometric polynomials  and they do
499: not change the general, polynomial form  of Eq.~\ref{eq:oneoverDgeneral}.
500: However, the second scaling function is not a polynomial with respect to
501: $\cos{f_i}$ or $\sin{f_i}$.  Yet Eq.~\ref{eq:oneoverD} involves a factor 
502: $\left(1 + e_j \cos{f_j}\right)$. It cancels out one power of $\left(1 + e_j
503: \cos{f_j}\right)$ appearing in the denominator of Eq.~\ref{eq:ij}. In order to
504: calculate the expansion in Eq.~\ref{eq:oneoverD}, for $l>0$ we differentiate
505: ${\cal D}(\alpha_{i,j})$  with respect to $\alpha_{i,j}$ as the composite
506: function (Eq.~\ref{eq:eqf}). This operation emerges factors  of the type of $A^r
507: B^s$, where $n = r + s \geq 1$. Looking at the  general form of $A$ and $B$ we
508: see that the term $\left(1 + e_j \cos{f_j}\right)$  appears with  natural powers
509: larger than $1$ and it cancels out  remaining $\left(1 + e_j \cos{f_j}\right)$
510: in the denominator of Eq.~\ref{eq:ij}. In this way, the general form of
511: trigonometric polynomial in Eq.~\ref{eq:oneoverDgeneral} is preserved. Still,
512: the free term in the Taylor expansion of ${\cal D}$ leads to an expression
513: involving $\left(1 + e_j \cos{f_j}\right)^{-1}$. Fortunately, we must integrate
514: such term with limits from $0$ to $2\pi$ and this effectively can be reduced to
515: averaging out $r_j$ over the orbital period (or the 
516: whole range of the true anomaly).
517: 
518: Finally, we can integrate Eq.~\ref{eq:rsecl} term by term.  Basically, the
519: problem has been reduced to  the calculation of definite integrals from products
520: of trigonometric functions $\sin(x)$ and $\cos(x)$ in some natural powers. These
521: integrals can be derived quite easily, at least in principle, nevertheless with
522: increasing order of the Taylor expansion, the calculations become extremely
523: tedious. To accomplish them, we used MATHEMATICA and fast AMD-Opteron computer.
524: 
525: The final result of the averaging is an expansion of the secular, two-body
526: Hamiltonian for a chosen pair of planets $i$ and $j$:
527: \begin{eqnarray}
528: & & \Hsec^{(i,j)} = -\frac{k^2 m_i m_j}{a_j} \times  \nonumber \\
529: && \quad \times \left[1 + \sqrt{1-e_j^2} \sum_{l=2}^{\infty}
530: {\left(\frac{\alpha_{i,j}}{1-e_j^2}\right)^l
531: \mathcal{R}^{(i,j)}_l(e_i,e_j,\Delta\varpi_{i,j})}\right].
532: \label{expansion}
533: \end{eqnarray}
534: Looking at the general form of this secular Hamiltonian, we learn that the role
535: of a formal parameter in the power-series  expansion of $\Hsec$ plays
536: (apparently) the following expression: 
537: \begin{equation} X_{i,j} \equiv \frac{\alpha_{i,j}}{1-e_j^2} =
538: \frac{a_i}{a_j (1-e_j^2)}. 
539: \end{equation} 
540: Obviously, these series cannot
541: converge if $X_{i,j} \geq 1$, hence we require that $X_{i,j}<1$.  Of course,
542: this is only the necessary (and as we see below, very rough) condition for the
543: convergence of these series. 
544: \rev{
545: However, as we explain in Sect.~2.5, attributing to $X_{i,j}$ 
546: the role of a 
547: parameter deciding on the convergence of these series is in fact misleading
548: because their divergence flows from quite a different source.
549: }
550: 
551: A few first terms of the expansion of the secular Hamiltonian,
552: Eq.~\ref{expansion}, are listed below. The free term (for $l=0$) is constant
553: because it depends on the {\em mean} semi-major axis $a_j$ only. The term with
554: $l=1$ vanishes identically.
555: 
556: Terms of order $2$ and $3$ may be identified with the quadrupole and octupole
557: secular Hamiltonian, respectively \citep{Ford2000,Lee2003}:
558: \begin{eqnarray}
559: \label{quadropole} 
560: & & \R^{(i,j)}_2 = \frac{1}{8} \left(3 e_i^2+2\right), \\
561: \label{octupole}
562: & & \R^{(i,j)}_3 = -\frac{15}{64}
563:     \left(3 e_i^2+4\right) e_i e_j \cos({\Delta{\varpi_{i,j}}}).
564: \end{eqnarray}
565: Higher order terms are the following:
566: \begin{eqnarray}
567: && \R^{(i,j)}_4 = \frac{9}{1024}
568:     \Big[ 
569:     70 \left(e_i^2+2\right) e_i^2 e_j^2 \cos({2\Delta{\varpi_{i,j}}})+\\
570:     && \quad +\left(15 e_i^4+40 e_i^2+8\right) \left(3 e_j^2+2\right)
571:     \Big],\nonumber \\
572: % \end{eqnarray}
573: %\begin{eqnarray}
574: && \R^{(i,j)}_5 = -\frac{105}{4096}
575:     \Big[
576:     7 \left(3 e_i^2+8\right)
577:     e_j^3 e_i^3 \cos({3\Delta{\varpi_{i,j}}}) +\\
578:     && \quad +2 \left(5
579:     \left(e_i^2+4\right) e_i^2+8\right) \left(3
580:     e_j^2+4\right) e_i e_j \cos({\Delta{\varpi_{i,j}}})
581:     \Big],\nonumber\\
582: && \R^{(i,j)}_6 = \frac{5}{65536} 
583:     \Big[
584:     2079 \left(3 e_i^2+10\right) e_i^4 e_j^4
585:     \cos({4\Delta{\varpi_{i,j}}}) +\\
586:     && \quad +630 \left(15 e_i^4+80
587:     e_i^2+48\right) \left(e_j^2+2\right) e_i^2 e_j^2 
588:     \cos({2\Delta{\varpi_{i,j}}}) +\nonumber\\
589:     && \quad +10 \left(35 e_i^6+210 e_i^4+168
590:     e_i^2+16\right) \left(15 e_j^4+40
591:     e_j^2+8\right)\Big].\nonumber
592: \end{eqnarray}
593: 
594: We computed the expansion up to the order of 24. This expansion is available on
595: the request in the form of raw MATHEMATICA input file; also available in
596: the form of on-line material after publishing this paper.
597: 
598: 
599: %_______________________________________________________________________________
600: \subsection{A comparison with the octupole theory}
601: %_______________________________________________________________________________
602: Here,  we compare the results provided by the secular theory derived in the
603: previous section with the results obtained with  the help of the  octupole
604: theory of two planets \citep{Lee2003} which has been obtained through averaging
605: the perturbation Hamiltonian with the help of von~Zeipel method up to the third
606: order in $\alpha_{1,2}\equiv \alpha$.  It has been applied to study qualitative
607: features of the secular dynamics in hierarchical planetary systems (i.e. with
608: small $\alpha$). A similar theory has been developed by \cite{Ford2000} who
609: investigated secular dynamics in  hierarchical triple stellar systems with large
610: separation of the third body. 
611: 
612: To make our discussion  more transparent, we use, in this section, the notation
613: of \cite{Lee2003}.  Their equations~(13), (14) and (15) have the following form:
614: \begin{eqnarray}
615: && L_1 = \frac{m_0 m_1}{m_0 + m_1} \sqrt{k^2 (m_0 + m_1) a_1},\\
616: && L_2 = \frac{(m_0 + m_1) m_2}{m_0 + m_1 + m_2} \sqrt{k^2 (m_0 + m_1 + m_2) a_2},\\
617: && G_j = L_j \sqrt{1 - e_j^2},
618: \end{eqnarray}
619: where $m_{1,2}$ are planetary masses, $m_0$ is the mass of the star, $j=1,2$
620: (we set the gravitational constant to $k^2$).  To derive the octupole theory in
621: terms of Jacobi reference frame, we start from writing down Eq.~\ref{Hfull} with
622: respect to Jacobi coordinates $\vec{r}_{1,2}$ of two point masses, as a sum of
623: two Keplerian terms and the perturbation:
624: \[
625: {\cal H}^J = \frac{1}{2\mu_1} \vec{p}^2_1
626:   -\frac{k^2 m_0 m_1}{\|\vec{r}_1\|}
627:   + \frac{1}{2\mu_2} \vec{p}_2^2 
628:     -\frac{k^2 m_0 m_2}{\|\vec{r}_2\|} + {{\HpertJ}},
629: \]
630: where
631: \[
632: {\HpertJ} = 
633:     -\frac{k^2 m_1 m_2}{\|\vec{r}_2 - (1-\kappa_1) \vec{r}_1\|}
634:     + k^2 m_0 m_2\left[ \frac{1}{\|\vec{r}_2\|} - 
635:     \frac{1}{\|\vec{r}_2 + \kappa_1 \vec{r}_1\|}
636:     \right].
637: \]
638: Here, $\vec{p}_{1,2}$ are the conjugate Jacobi momenta,
639: $\kappa_1=m_1/(m_0+m_1)$, and the reduced
640: masses are
641: \[
642: \mu_1 = \frac{m_1 m_0}{(m_0+m_1)},
643:  \quad \mu_2=\frac{m_2 (m_0+m_1)}{(m_0+m_1+m_2)}.
644: \]
645: For details, see, e.g., \citep{Malhotra1993}. Now,  after expanding ${\HpertJ}$
646: with respect to small $\kappa_{1}$ and retaining first order terms, we can
647: show that the  perturbation  has the same form as Eq.~\ref{Hpert} with the
648: accuracy to the second order in the masses $m_{1,2}/m_0$
649: \citep{Malhotra1993}:
650: \[
651: \Hint = - k^2 m_1 m_2 \left[ \frac{1}{\|\vec{r}_2-\vec{r}_1\|}
652:  - \frac{\vec{r}_1 \cdot \vec{r}_2}{\|\vec{r}_2\|^3} \right].
653: \] 
654: The same truncated Hamiltonian is analyzed in \citep{Libert2005} who derived the
655: secular Hamiltonian of the 12-th order in eccentricities by the "averaging with
656: scissors" (i.e., by eliminating  from the Fourier expansion of $\Hint$ all fast
657: periodic terms dependent on $l_i$). These authors report that their secular
658: theory reproduces qualitatively results of \cite{Michtchenko2004} on the
659: analytical way. 
660: 
661: The indirect part of the truncated Hamiltonian $\Hint$ also averages out to a
662: constant, hence the rest of the averaging process is the same as  in the case of
663: $\Hpert$ written with respect to  the Poincar\'e elements. Nevertheless, the
664: averaged $\Hint$ is missing  terms of orders higher than two in the planetary 
665: masses. Indeed, with our method we derived the secular octupole Hamiltonian
666: which has the same functional form as formulae (17) in \citep{Lee2003}. 
667: However, there are some differences in coefficients $C_2$ and $C_3$ [see their
668: equations (18) and (19)].  These coefficients in our expansion are the
669: following:
670: \begin{eqnarray}
671: && C_2 = \frac{1}{16} \frac{G^2 (m_0 + m_1)^7 m_2^7}{(m_0 + m_1 + m_2)^3 (m_0 m_1)^3}
672: \frac{L_1^4}{L_2^3 G_2^3} D_2,\\
673: && C_3 = \frac{15}{64} \frac{G^2 (m_0 + m_1)^9 m_2^9 (m_0 - m_1)}{(m_0 + m_1 + m_2)^4 (m_0
674: m_1)^5} \frac{L_1^6}{L_2^3 G_2^5} D_3,
675: \end{eqnarray}
676: where we extracted out two factors leading to the  difference between the
677: respective formulae:
678: \begin{eqnarray}
679: && D_2 = \frac{m_0 + m_1}{m_0} \sim 1 + O\left(\frac{m_1}{m_0}\right), \\
680: && D_3 = \frac{(m_0 + m_1)^2}{m_0 (m_0 - m_1)} \sim 1 +
681: O\left(\frac{m_1}{m_0}\right).
682: \end{eqnarray}
683: These factors can be thought as equal to $1$ in \citep{Lee2003}. However, the
684: theories are consistent within the assumed accuracy of the expansion and  the
685: relative magnitude of terms skipped from ${\HpertJ}$ [of the order of
686: $O(m_1/m_0)$].
687: 
688: Actually, our averaging algorithm can be applied also to the {\em full}
689: perturbing Hamiltonian, thanks to straightforward generalization for  terms like
690: the following: 
691: \[
692: \frac{1}{\| \delta_1 \vec{r}_1 - \delta_2 \vec{r}_2\|},
693: \]
694: where $\delta_1,\delta_2$ are some constants. In that instance, we obtain exactly the
695: same $C_{2,3}$ as in \citep{Lee2003}. Hence,  as one would expect, both
696: approaches lead to fully equivalent results.
697: 
698: This comparison also reveals  that the averaging of  the truncated Hamiltonian
699: is in fact quite problematic  because the accuracy of the secular expansion has
700: nothing to do with the magnitude of the rejected terms. Already for Jupiter-mass
701: planets, a contribution of these terms may be significant (see Sect.~2.4 for
702: details).
703: 
704: Moreover, a direct comparison of the theories would be  more subtle. Although
705: the functional forms of the perturbing Hamiltonians are the same,  they are
706: expressed in terms of two different sets of canonical variables. Hence, the mean
707: elements $a, e, \varpi$ have different meaning in these theories, i.e., the same
708: physical configuration of the planets will be parameterized with quantitatively
709: different values of the mean elements.
710: 
711: %_______________________________________________________________________________
712: \subsection{Tests of the analytic secular theory}
713: %_______________________________________________________________________________
714: \label{sec:tests}
715: To test the accuracy and relevance of the high order expansion of $\Hsec$, we
716: calculated the  magnitude of  subsequent terms  relative to the free term. This
717: expansion is computed up to the order of 24 for parameters
718: ($\mu_{i,j},\alpha_{i,j}$) of extrasolar planetary systems taken from  the Jean
719: Schneider  Encyclopedia of Extrasolar Planets\footnote{http://exoplanets.eu}. 
720: The results of this experiment are presented in Fig.~\ref{accuracy1}. Each panel
721: in this figure is labeled with the name of a relevant star and a number of
722: putative planets it hosts (written in brackets).  As we can see, for all 
723: examined systems, the sum of terms in $\Hsec$  of the same order (note that we
724: can have two and more planets in the system) decrease rapidly with the order of
725: the expansion. For a few most separated systems  with two planets (e.g.,
726: HD~217107, HD~190360), the highest order terms are as low as $\sim 10^{-37}$ in 
727: the relative magnitude.  In the tested sample, the largest 24th-order terms are
728: $\sim 10^{-7}$.  Hence, the secular energy can be  calculated with excellent
729: accuracy.  We note that similar tests of the precision of the secular theory were
730: presented in \citep{Gallardo2005} and \citep{Libert2005}.
731: \begin{figure*}
732: \centerline{
733: \includegraphics [width=17.6cm]{fig1.eps}
734: }
735: \caption{
736: Convergence of the expansion of the $N$-planet secular Hamiltonian derived in
737: this paper.  Each panel is for the magnitude of expansion terms
738: in~Eq.~\ref{expansion} divided by free term. The $x$-axis is for the order  of
739: the expansion, the $y$-axis is for  $\log_{10}\,\|{\cal H}_n\|$, where ${\cal
740: H}_n$ denotes the magnitude of the sum of $n$-th order terms divided by the  sum
741: of the free terms.  Parameters $(\mu_{i,j},\alpha_{i,j})$ of $\Hsec$  are chosen
742: for the known multi-planet extrasolar planetary systems. Their parent stars are
743: marked in each panel together with number of planets written in brackets.
744: }
745: \label{accuracy1}
746: \end{figure*}
747: 
748: In the second test,  we compare the outcome of our algorithm with the results of
749: semi-analytical approach  by \cite{Michtchenko2004,Michtchenko2006} [see also
750: \cite{Migaszewski2008} for some technical aspects] who applied the method to
751: study the secular dynamics of two-planet system. In the algorithm of
752: \cite{Michtchenko2004} the perturbing Hamiltonian is averaged out by means of
753: the numerical integration. Hence, one avoids any expansion of the Hamiltonian
754: and the results are formally exact (or very accurate, providing that precise
755: enough quadratures are applied). 
756: 
757: To examine the  accuracy of the analytic secular theory, we calculated the
758: levels of the secular Hamiltonians for a number of extrasolar planetary systems
759: which can be potentially regarded as non-resonant or well fitting assumptions of
760: the secular theory. Because the phase space of the secular problem is
761: three-dimensional [i.e., $(G_1,G_2,\Delta\varpi)$], we choose the so called
762: representative plane of the initial conditions to plot the  secular energy
763: levels. The definition of the representative plane follows
764: \cite{Michtchenko2004}. The secular Hamiltonian of a coplanar two-planet system
765: depends only on $\Delta\varpi=\varpi_2-\varpi_1$ and eccentricities $(e_1,e_2)$
766: coupled through the integral of the total angular momentum   (or  the so called 
767: Angular Momentum Deficit, $AMD = G_1 + G_2$). Effectively, the secular problem
768: has one degree of freedom and is integrable.  \cite{Michtchenko2004} have shown
769: that {\em all} phase trajectories of the non-resonant system pass through a
770: plane defined with $\Delta\varpi = 0$ or $\Delta\varpi = \pi$ [see also
771: \citep{Pauwels1983} for the qualitative analysis of the secular two-planet
772: problem]. The representative plane may be defined for fixed $\alpha_{1,2} \equiv
773: \alpha = a_1/a_2$ and $\mu_{1,2} \equiv\mu=m_1/m_2$ as follows:
774: \[
775: {\cal S} = \{e_1 \cos\Delta\varpi  \times e_2\, ;\ 
776: e_1 \in [0,1),\, e_2 \in [0,1), \, \Delta\varpi=0 \, \cup \, \Delta\varpi = \pi
777: \}.
778: \]
779: This plane comprises of two $(x=e_1 \cos\Delta\varpi, y=e_2)$-half-planes with
780: $x \leq 0$ for $\Delta\varpi = \pi$ and with  $ x \geq 0$ for $\Delta\varpi =0$.
781: Simultaneously, the derivatives of the secular Hamiltonian with respect to
782: $\Delta\varpi$ are equal to zero for $\Delta\varpi=0,\pi$.   It follows from the
783: symmetry of interacting orbits with respect to both apsidal lines.  Having  the
784: secular Hamiltonian in explicit analytic form, we can verify this property 
785: directly.  Indeed, each term in the secular Hamiltonian depends on
786: $\Delta\varpi$ only through $\cos(l \Delta\varpi)$ with $l \in {\mathbb N}$,
787: $l>0$ and it means that $\Hsec$ is even function of $\Delta\varpi$. Hence,
788: $\partial\, \Hsec/\partial\,\Delta\varpi$ implies factors involving $\sin(l
789: \Delta\varpi)$ and these terms vanish identically for  $\Delta\varpi = 0,\pi$. 
790: { Formally, it is  possible  that $\partial\, \Hsec/\partial\,\Delta\varpi=0$
791: also for $\Delta\varpi \neq 0,\pi$, however, to find such solutions we should
792: solve highly nonlinear equation involving $(e_1,e_2)$ and trigonometric
793: functions of $\Delta\varpi$.}
794: 
795: Each pair of $(e_1, e_2)$ for which $\partial\, \Hsec/\partial\,G_1=0$
796: corresponds to an equilibrium in the secular problem (simultaneously, they are
797: the extrema of the secular Hamiltonian).  These equilibria appear both in the
798: negative half-plane of ${\cal S}$, as mode~II solutions \rev{(this mode is
799: Lyapunov stable,  and may be characterized with librations of angle 
800: $\Delta\varpi$ around $\pi$ in the  evolution of neighboring orbits}), and in
801: the positive half-plane of ${\cal S}$ as mode~I solutions 
802: \rev{(Lyapunov stable, with librations of angle $\Delta\varpi$ around $0$
803: of the nearby orbits).
804: } 
805: Further, in the regime of large eccentricities, in the positive half-plane a
806: new, non-classic mode of motion may appear [it is the so called non-linear
807: secular resonance,  NSR from hereafter, see \citep{Michtchenko2004} for
808: details]. These results are derived through the numerical (exact) approach,
809: hence their reproduction by the analytical theory provides an absolute test of
810: its quality and accuracy.
811: 
812: \begin{figure*}
813: \centerline{
814: \vbox{
815:     \hbox{\includegraphics [width=58mm]{fig2a.eps}\hskip-3mm
816:           \includegraphics [width=58mm]{fig2b.eps}\hskip-3mm
817:           \includegraphics [width=58mm]{fig2c.eps}}
818:     \hbox{\includegraphics [width=58mm]{fig2d.eps}\hskip-3mm
819:           \includegraphics [width=58mm]{fig2e.eps}\hskip-3mm
820:           \includegraphics [width=58mm]{fig2f.eps}}
821:     \hbox{\includegraphics [width=58mm]{fig2g.eps}\hskip-3mm
822:           \includegraphics [width=58mm]{fig2h.eps}\hskip-3mm
823:           \includegraphics [width=58mm]{fig2i.eps}}
824: }
825: }
826:  \caption{
827: A test of the secular theory of a coplanar two-planet system, derived in this
828: paper. Each panel is for the representative energy plane,   $(e_1
829: \cos{\Delta{\varpi}}, e_2)$.  In the right half-plane, $\Delta{\varpi}=0$,  at
830: the left half-plane, $\Delta{\varpi}=\pi$.   Gray region corresponds to crossing
831: orbits and its boundary is defined through $a_2 (1 - e_2) = a_1 (1 -  e_1)$,
832: where (1) is for the inner orbit and (2) is for the outer orbit. Black
833: thick line marks the ``anti-collision'' line defined with $a_2 (1 - e_2) = a_1
834: (1 + e_1)$ (see the text for more details). Black, thin lines are for  contour levels of  the last two terms
835: of expansion Eq.~\ref{expansion}, i.e., the sum of absolute values of the
836: $23$-th and $24$-th order terms divided by the free term.   A few contour
837: levels  ($1$, $10^{-8}$, $10^{-16}$ and $10^{-24}$, respectively) are  marked
838: with  thicker curves. The area colored in orange determines the region where the
839: expansion of $\Hsec$ is divergent. Thick blue  lines mark the positions of
840: stable equilibria: mode~II (with apsides anti-aligned, the left half-plane of
841: the representative plane), and mode~I (with apsides  aligned, the right-half
842: plane).  These solutions are obtained numerically with the help of
843: semi-analytical averaging algorithm in \citep{Michtchenko2004}. Green curves are
844: for the nonlinear-secular resonance (NSR).  Red curves are for the corresponding
845: libration modes  calculated analytically with the help of the 24-order expansion
846: of $\Hsec$. Labels in the top-left corner at each panel are for the parameters
847: of the expansion, $\alpha \equiv a_1/a_2$  and $\mu \equiv m_1/m_2$.
848: }
849: \label{accuracy2}	  
850: \end{figure*}
851: %
852: The results of  the second experiment are presented in Figure~\ref{accuracy2}. 
853: Each panel in this figure is for the representative plane of initial conditions
854: computed and calculated for two-planet systems characterized with mass ratio
855: $\mu$ and semi-major axes ratio $\alpha$. These parameters are  written in the
856: top-left corner at each relevant panel. To illustrate the significance of
857: highest order terms in Eq.~\ref{expansion}, we plot contour levels of  
858: $(\|{\cal H}_{23}\| + \|{\cal H}_{24}\|)/\|{\cal H}_0\|$, i.e.,  the relative
859: magnitude of the sum of the last two terms with respect to the magnitude of the
860: free term (in general, ${\cal H}_n$ would stand for a sum of expansion terms
861: over the number of planet pairs in the given multi-body configuration). Four
862: particular contour levels of $1$, $10^{-8}$, $10^{-16}$  and $10^{-24}$,
863: respectively, are distinguished with thicker lines and labeled accordingly.
864: 
865: In each panel, we also marked positions of stationary solutions (modes I, II,
866: and~NSR).  Solutions represented by  thick curves are derived with the help of
867: numerical averaging \citep{Michtchenko2004}: blue curves are for stable
868: equilibria, and  thick green curves are for unstable solutions (the {\em
869: non-linear secular resonance}). These curves represent practically exact (or
870: very precise) solution to the problem. The thin, red lines mark the positions of
871: equilibria calculated analytically, with the help of the $24$th-order expansion
872: of $\Hsec$.
873: 
874: The results can be summarized with a few  interesting conclusions. Clearly, in
875: the regions of the representative  $(e_1\cos\Delta\varpi,e_2)$-plane at which 
876: $(\|{\cal H}_{23}\| + \|{\cal H}_{24}\|)/\|{\cal H}_0\| < 10^{-3}$, the
877: precision of the analytical method is excellent. The secular theory predicts
878: exactly positions of the equilibria in the negative half-plane of ${\cal S}$
879: (when $\Delta{\varpi} = \pi$) no matter how large is $\alpha$. On contrary, the
880: exact derivation of the shape of the non-linear secular resonance is a
881: challenging problem for the analytic approach \citep[see also][]{Henrard2005}.
882: The nonlinear resonance can be reproduced well by the analytic theory providing
883: that $(\|{\cal H}_{23}\| + \|{\cal H}_{24}\|)/\|{\cal H}_0\| < 10^{-3}$. This is
884: also an empirical border of the convergence of the secular expansion.  We also
885: found another empirical convergence condition that follows from the notion of
886: the geometric series, i.e,
887: $
888: \|{\cal H}_{24}/{\cal H}_{22}\|<1.
889: $
890: This inequality is illustrated with triangular, orange colored regions labeled
891: with ``$>1$''. In these regions, the series are divergent, hence the secular
892: theory cannot reproduce the {\em real} dynamics.  This may be interpreted as a
893: clear limitation of the analytic theory. In fact, as we show below (Sect.~2.5),
894: this problem appears rather due to imperfect algorithm of the expansion,  which
895: can be still improved. On the other hand, the results of this test  provide an
896: example that illustrates excellent properties of the semi-analytical  approach
897: invented by \cite{Michtchenko2004}.
898: 
899: {
900: %_______________________________________________________________________________
901: \subsection{An improved averaging algorithm}
902: %_______________________________________________________________________________
903: A real source of the divergence of the secular series, Eq.~\ref{expansion}, can
904: be deduced after we draw in Fig.~\ref{accuracy2} the ``anti-collision'' line 
905: defined through $a_2 (1 - e_2) = a_1 (1 + e_1)$ (see the right-half of the
906: representative plane). Clearly, the series diverge above this line  and the
907: positions of the equilibria are strongly distorted. In this area, for some
908: points or parts of the orbits, $r_i[E_i(t)]>r_j[f_j(t)]$, while in the expansion
909: Eq.~\ref{expansion},  $r_i<r_j$ {\em must} be satisfied in the
910: whole ranges of the anomalies.  That may  happen when the pericenter of
911: the outer orbit is closer to the star than the apocenter of the inner orbit, no
912: matter what is the relative orientation of their
913: apsidal lines. Then the condition of $r_i<r_j$, which required to write down
914: Eq.~\ref{eq:squareroot2}, is violated. Obviously, it may be expressed by the
915: equation of the anti-collision line, and in other
916: words, through the requirement that the inner orbit lies inside a circle of a radius equal to
917: the pericenter distance of the outer orbit. Now it is also clear why for the
918: apsides anti-aligned, we have always very good convergence of the secular series
919: while in the right half-plane, the convergence region is generally 
920: strongly limited. Hence, the convergence limit of the expansion described in Sect.~2 may
921: be simply interpreted through the conditions for the collision lines.
922: We may also note that the problem persists in any secular theory that relies
923: directly on Eq.~\ref{eq:squareroot2}. One should be also aware that the
924: conditions of crossing orbits do not depend on masses.
925: If the masses are large, the  dynamical collision curve
926: appears for much smaller eccentricity than the geometrical collision line.
927: \citep[e.g.,][]{Gozdziewski2007,Michtchenko2008}.
928: 
929: A cure for the divergence problem  may be a modified expansion of the term
930: $1/\Delta_{i,j}$, helping us to construct  a secular  theory that has no limits
931: of the type $a_i (1 + e_i) < a_j (1 - e_j)$. To derive the secular  expansion in
932: Sect.~2, at first we factor $r_j$ in Eqs. (\ref{squareroot1}), 
933: (\ref{eq:squareroot2}). To have the series convergent for all positions of
934: planets on their orbits, we propose to factor from the square
935: root~(\ref{squareroot1}) a term consisting of $a_2$ multiplied by some scale
936: factor, $\eta \geq 1$ (instead of $r_j$). Then we can express the distance
937: between planets $i$ and $j$ as follows:
938: \begin{equation}
939: \Delta_{i,j} = \eta a_j \sqrt{1 + \zeta_{i,j} (\vec{r}_i,\vec{r}_j)},
940: \label{newexp}
941: \end{equation}
942: where
943: \begin{equation}
944: \zeta_{i,j} = \frac{1}{\eta^2 a_j^2} \left[r_i^2 + r_j^2 - 2 \vec{r}_i \cdot \vec{r}_j - \eta^2
945: a_j^2\right] \equiv 
946:  \frac{1}{\eta^2 a_j^2} \left[\Delta_{i,j}^2 - \eta^2 a_j^2\right].
947: \label{zeta}
948: \end{equation}
949: The requirement of the convergent expansion of $\Delta_{i,j}^{-1}$ with respect
950: to $\zeta_{i,j}$  implies $\|{\zeta}_{i,j}\|<1$. If $\|\zeta_{i,j}\|=1$  the
951: distance between planets is equal to $\sqrt{2} \eta a_j$. The convergence
952: condition is fulfilled when   $\max \Delta_{i,j} < \sqrt{2} \eta a_j$ for the
953: given orbits. The maximal distance between planets in coplanar orbits may be
954: bounded by $a_i (1 + e_i) + a_j (1 + e_j)$. Then  the factor $\eta$ has the 
955: form of:
956: \begin{equation}
957: \eta = \frac{a_i (1 + e_i) + a_j (1 + e_j)}{\sqrt{2} a_j} = \frac{1}{\sqrt{2}}
958: \left[\alpha_{i,j} (1 + e_i) + (1 + e_j)\right].
959: \label{eta}
960: \end{equation}
961: For hierarchical systems with small eccentricity of the outer planet, $\eta
962: \approx 1$. For more compact systems with  large eccentricities, $\eta \approx
963: \sqrt{2}$. For non-coplanar systems, $\eta$ may be  even larger. In
964: practice, this parameter  makes it possible to control the convergence rate of
965: the secular expansion. The convergence rate will be faster for large distances
966: $\Delta$ but slower for smaller distances, moreover the condition of
967: $-1<\zeta_{i,j}<1$ should be always fulfilled.
968: 
969: The term $1/\Delta_{i,j}$ may be expanded with respect to $\zeta_{i,j}$:
970: \begin{equation}
971: \frac{1}{\Delta_{i,j}} = \frac{1}{\eta a_j}
972: \left[ 1 +  \sum_{l=1}^{\infty} {\frac{(-1)^l (2 l - 1)!!}{2^l l!} 
973: \zeta_{i,j}^l} \right].
974: \label{exp2}
975: \end{equation}
976: To average out the above formulae, we express positions of both planets in a
977: given pair with respect to  the eccentric anomaly. Next, we change the
978: integration variables similarly as in Sect.~2, i.e.,
979: $
980: d\,M_k = I_k(E_k, e_k) d\,E_k.
981: $ 
982: We also should express $\vec{r}_i \cdot \vec{r}_j$ through $\Delta{\varpi}$
983: (again, fixing the reference frame with the apsidal line of the inner orbit).
984: After expressing terms $r_i, r_j, \vec{r}_i \cdot \vec{r}_j$ through eccentric
985: anomalies, the  function $\zeta \equiv \zeta_{1,2}$ has the
986: following explicit form
987: (to shorten the notation, let us fix $i\equiv 1$, $j \equiv 2$ for a
988: given pair of planets):
989: :
990: \begin{eqnarray}
991: &&\zeta = \frac{1}{\eta ^2} \Big[\theta_0 + \theta_1 \cos{E_1} + \theta_2 \cos{E_2} +
992: \theta_3 \sin{E_1} + \\ 
993: && + \theta_4 \sin{E_2} +
994: \theta_5 \cos{E_1}^2 + \theta_6 \cos{E_2}^2 + \theta_7 \cos{E_1} \cos{E_2} + \nonumber\\
995: && + \theta_8 \sin{E_1}
996: \sin{E_2} + \theta_9 \cos{E_1} \sin{E_2} + \theta_{10} \cos{E_2} \sin{E_1}\Big],\nonumber
997: \end{eqnarray}
998: where ${E}_1, E_2$ are eccentric anomalies of the inner and outer planets
999: respectively, and coefficients
1000: $\theta_l, l \geq 0$, read as follows:
1001: \begin{eqnarray}
1002: && \theta_0 = \alpha^2-2 \alpha e_1 e_2 \cos{\Delta{\varpi}}+1-\eta^2,\nonumber\\
1003: && \theta_1 = 2 \alpha e_2 \cos{\Delta{\varpi}}-2 \alpha^2 e_1,\nonumber \\
1004: && \theta_2 = 2 \alpha e_1 \cos{\Delta{\varpi}}-2 e_2,\nonumber \\
1005: && \theta_3 = -2 \alpha e_2 \sqrt{1-e_1^2} \sin{\Delta{\varpi}},\nonumber \\
1006: && \theta_4 = 2 \alpha e_1 \sqrt{1-e_2^2} \sin{\Delta{\varpi}},\nonumber \\
1007: && \theta_5 = \alpha^2 e_1^2, \\
1008: && \theta_6 = e_2^2, \nonumber \\
1009: && \theta_7 = -2 \alpha \cos{\Delta{\varpi}}, \nonumber \\
1010: && \theta_8 = -2 \alpha \sqrt{1-e_1^2} \sqrt{1-e_2^2} \cos{\Delta{\varpi}},\nonumber\\
1011: && \theta_9 = -2 \alpha \sqrt{1-e_2^2} \sin{\Delta{\varpi}},\nonumber \\
1012: && \theta_{10} = 2 \alpha \sqrt{1-e_1^2} \sin{\Delta{\varpi}}.\nonumber
1013: \end{eqnarray}
1014: Here, $\alpha \equiv \alpha_{1,2}$, $\Delta{\varpi} \equiv
1015: \Delta{\varpi_{1,2}}$  and $e_1, e_2$ are  the eccentricities of the inner and
1016: outer planet, respectively.
1017: 
1018: After the double averaging of $\zeta$ over the mean anomalies we 
1019: obtain:
1020: \begin{equation}
1021: <\zeta> = \frac{1}{2 \eta^2} \bigg[\left(3 e_1^2+2\right) \alpha^2-9 \alpha
1022:     e_1 e_2 \cos{\Delta{\varpi}}
1023:     +3 e_2^2-2 \eta^2+2\bigg].\nonumber
1024: \end{equation}
1025: The averaging of the square of $\zeta$ over  the mean anomalies brings the
1026: following formulae:
1027: \begin{eqnarray}
1028: &&<\zeta^2> = \frac{1}{8 \eta^4} 
1029:     \bigg[
1030:     \alpha^4 \left(8 + 40 e_1^2 + 15 e_1^4\right) + \nonumber\\
1031: &&  + \alpha^3 \left(-30 e_1 e_2 (3 e_1^2 + 4) \cos{\Delta{\varpi}}\right) + \\
1032: &&  + \alpha^2 \left[(2 - \eta^2) (16 + 24 e_1^2) + e_1^2 e_2^2 (72 + 100 \cos{2\Delta{\varpi}})
1033: + 48 e_2^2 \right] + \nonumber\\
1034: && + \alpha \left(-6 e_1 e_2 (20 - 12 \eta^2 + 15 e_2^2) \cos{\Delta{\varpi}}\right) + \nonumber\\
1035: && + 8 (\eta^2 - 1)^2 + 40 e_2^2 - 24 e_2^2 \eta^2 + 15 e_2^4
1036:     \bigg].\nonumber
1037: \end{eqnarray}
1038: These preliminary calculations show that the new algorithm leads to more complex
1039: expansion of the secular Hamiltonian than the simple approach in Sect.~2.2
1040: which, as we have demonstrated, is limited in some cases. Moreover, we found
1041: this improvement after submitting the manuscript, hence the new expansion and  a
1042: detailed study of its properties would make the paper very lengthy.   We are
1043: going to present the improved algorithm and the results of its tests in a new
1044: work devoted to the analytic theory of non-coplanar model of $N$-planets. A
1045: generalization of Eq.~\ref{eta} for that case seem straightforward, because we
1046: should only calculate $\vec{r}_i \cdot \vec{r}_j$ with the help of appropriate
1047: rotation matrix parameterized through Euler angles, i.e., the Keplerian elements
1048: ($i_i,i_j,\omega_i,\omega_j,\Omega_i,\Omega_j$). Hence, only the coefficients
1049: $\theta_l$ will be modified.
1050: }
1051: 
1052: %_______________________________________________________________________________
1053: \section{Secular dynamics of three-planet system}
1054: %_______________________________________________________________________________
1055: The two-planet  secular Hamiltonian in Sect.~2 can be easily adapted to
1056: construct the secular theory for $N$-planet system.  At present, a few
1057: candidates of such configurations are already discovered, including four planet
1058: systems, e.g., $\mu$~Arae
1059: \citep{Jones2002,Butler2006,Gozdziewski2007a,Pepe2007},  five or even six planet
1060: configuration around 55~Cnc \citep{Fischer2007}.  Here, as the simplest and
1061: most  natural generalization of the two-planet model, we consider the secular
1062: theory of $three$-planet configuration which is far from MMRs
1063: and collision zones.
1064: 
1065: The three-planet model is described by Hamiltonian in Eqs.~(\ref{Hfull}),
1066: (\ref{HKepl}) and (\ref{Hpert}), respectively, where $N=3$.  Because the nodal
1067: longitudes are undefined in the coplanar system, and the secular dynamics
1068: depends on the relative positions of the {\em mean} orbits, we can eliminate the
1069: nodal longitudes from the problem. Let indices $i=1,2,3$ enumerate the planets.
1070: Their semi-major axes are $a_1 < a_2 < a_3$, respectively. After averaging
1071: $\Hfull$ over the mean anomalies, the secular Hamiltonian $\Hsec$ does not
1072: depend on $l_i$ anymore. Therefore, conjugate momenta $L_i$ (hence,  the
1073: semi-major axes) are constants of motion. Because the secular system does not
1074: depend on  particular longitudes of nodes, the respective degree of freedom is
1075: also irrelevant for the secular dynamics.
1076: 
1077: Hence, the secular  system can be described with the following set of canonical
1078: elements:
1079: \begin{eqnarray}
1080: && g_1 = -\varpi_1, \qquad G_1,  \nonumber\\
1081: && g_2 = -\varpi_2, \qquad G_2,          \\
1082: && g_3 = -\varpi_3, \qquad G_3. \nonumber
1083: \label{trans2}
1084: \end{eqnarray}
1085: We can eliminate one more degree of freedom with the help of the angular
1086: momentum  integral, which can be also expressed with $AMD = G_1 + G_2 + G_3$. 
1087: For that purpose, we perform the following canonical transformation:
1088: \begin{eqnarray}
1089: \label{trans3}
1090: && \sigma_1 = g_1 - g_3 \equiv \varpi_3 - \varpi_1 \equiv
1091: \Delta\varpi_{1,3},  \quad G_1, \nonumber\\
1092: && \sigma_2 = g_2 - g_3 \equiv \varpi_3 - \varpi_2 \equiv
1093: \Delta\varpi_{2,3},  \quad G_2, \\
1094: && \sigma_3 = g_3 \equiv -\varpi_3,  \quad AMD = G_1 + G_2 + G_3, \nonumber
1095: \end{eqnarray}
1096: introducing new canonical angles $\sigma_1,\sigma_2,\sigma_3$.  These angles 
1097: can be interpreted as two-planet $\Delta\varpi$ defined for each pair of planets
1098: in the three-planet system. The secular Hamiltonian can be expressed through
1099: $\sigma_1$ and $\sigma_2$ explicitly, hence $\sigma_3$ is cyclic and $AMD$ is
1100: constant of motion. Actually, we reduced the secular system of three planets to
1101: two degrees of freedom, with the secular energy and $AMD$ as free parameters.
1102: 
1103: %_______________________________________________________________________________
1104: \subsection{Representative planes of initial conditions}
1105: %_______________________________________________________________________________
1106: Now, we have a similar problem as in the case of two-planet configuration. We
1107: want to illustrate the dynamical properties of the system in possibly global
1108: manner. To characterize its  dynamical states, we follow the general idea of the
1109: representative plane of initial conditions. Here, we focus on the equilibria in
1110: the secular problem. Fixing the integral of $AMD$ as a parameter of the system,
1111: the dynamics may be represented in four-dimensional phase space of
1112: $(G_1,G_2,\sigma_1,\sigma_2)$, or $(e_1,e_2,\sigma_1,\sigma_2)$. The
1113: representative plane will be chosen according with:
1114: \begin{equation}
1115: \frac{\partial \Hsec}{\partial \sigma_1} = 0, \quad
1116: \quad \frac{\partial \Hsec}{\partial \sigma_2} = 0.
1117: \label{condition_non_axial}
1118: \end{equation}
1119: Then any pair of points $(e^0_1,e^0_2)$ that belongs to
1120: the representative plane, and the following equations
1121: are satisfied:
1122: \begin{equation}
1123: \frac{\partial \Hsec}{\partial G_1} = 0, \quad
1124: \quad \frac{\partial \Hsec}{\partial G_2} = 0,
1125: \label{conditionG}
1126: \end{equation}
1127: defines an equilibrium of the secular problem. Note that the last conditions
1128: implies also $\partial\,\Hsec/\partial\,G_3=0$ because $G_3 \equiv G_3(G_1,G_2)$
1129: is a function parameterized by the total angular momentum (or $AMD$).  The
1130: explicit and simple transformation between the eccentricity and the element~$G$
1131: makes it possible to solve the above conditions in terms of $(e_1,e_2)$.
1132: 
1133: Actually, the most obvious definition  of the representative plane is  a
1134: generalization of that plane constructed  in the two-planet problem. For fixed
1135: $a_1,a_2,a_3$, $m_1,m_2,m_3$ and $AMD$, as a parameter, the {\em symmetric}
1136: representative plane is the set of points such that
1137: \[
1138: {\cal S} = \{e_1 \cos\sigma_1  \times e_2 \cos\sigma_2 ;\ 
1139: e_1 \in [0,1),\, e_2 \in [0,1), \, \sigma_{1,2}=0 \cup \pi
1140: \}.
1141: \]
1142: This plane comprises of four quaters. The signs of $e_{1,2}$ of the  coordinated
1143: axes tell us on the respective values of the secular angles. In that case, the
1144: condition in Eq.~\ref{condition_non_axial} is fulfilled thanks to apsidal
1145: symmetries of the problem. It is also obvious by recalling that ${\Hsec}$ is
1146: even function of $\sigma_{1,2}$. The derivatives of ${\Hsec}$ over
1147: $\sigma_{1,2}$ depend on  factors involving $\sin (l\sigma_{1,2})$, $l \in
1148: {\mathbb N}$, $l>0$ that vanish for $\sigma_{1,2} = 0,\pi$.  Alternatively, the
1149: representative plane may be also defined through $\sin{\sigma_{1,2}}=0$.
1150: 
1151: Moreover the condition for the representative plane, defined through  the
1152: vanishing derivatives over secular angles, may be also fulfilled for
1153: $\sin{\sigma_{1,2}} \neq 0$. Let us start with the octupole (third-order)
1154: approximation of the secular Hamiltonian. Using Eqs.~(\ref{expansion}),
1155: (\ref{quadropole}) and (\ref{octupole}),  we can write the secular Hamiltonian
1156: in the following short form of:
1157: \begin{equation}
1158: \Hsec = \gamma_{1,2} \cos{\Delta{\varpi}_{1,2}} + \gamma_{1,3} 
1159: \cos{\Delta{\varpi}_{1,3}} + \gamma_{2,3} 
1160: \cos{\Delta{\varpi}_{2,3}} + \gamma_{4}.
1161: \end{equation}
1162: Using the canonical angles defined with Eq.~\ref{trans3}:
1163: \begin{equation}
1164: \Hsec = \gamma_{1,2} \cos{(\sigma_1-\sigma_2)} + \gamma_{1,3} \cos{\sigma_1} + 
1165: \gamma_{2,3} \cos{\sigma_2} + \gamma_{4},
1166: \end{equation}
1167: where $\gamma_{1,2}, \gamma_{1,3}, \gamma_{2,3}, \gamma_4$ are functions of $e_1, e_2, e_3$.  
1168: Hence, the {\em non-symmetric representative plane} is defined through the following
1169: conditions, the same as Eq.~\ref{condition_non_axial}, 
1170: in the explicit form:
1171: \begin{eqnarray}
1172: -\gamma_{1,2} \sin{(\sigma_1-\sigma_2)} - \gamma_{1,3} \sin{\sigma_1}=0,\nonumber\\
1173: \gamma_{1,2} \sin{(\sigma_1-\sigma_2)}  - \gamma_{2,3} \sin{\sigma_2}=0.
1174: \label{representative_conditions}
1175: \end{eqnarray}
1176: Obviously, conditions in Eq.~\ref{representative_conditions} are satisfied not only
1177: when $\sin{\sigma_{1,2}} = 0$.  We have four other solutions, satisfying 
1178: Eq.~\ref{representative_conditions}, i.e.:
1179: \begin{eqnarray}
1180: \sigma_1 = \pm \arccos \left[-\frac{1}{2} \gamma_{1,2} \left(\frac{1}{\gamma_{2,3}} + 
1181: \frac{\gamma_{2,3}}{\gamma_{1,2}^2} - \frac{\gamma_{2,3}}{\gamma_{1,3}^2}\right)\right],\nonumber\\
1182: \sigma_2 = \pm \arccos \left[-\frac{1}{2} \gamma_{1,3} \left(\frac{1}{\gamma_{1,2}} 
1183: + \frac{\gamma_{1,2}}{\gamma_{1,3}^2} - \frac{\gamma_{1,2}}{\gamma_{2,3}^2}\right)\right].
1184: \end{eqnarray}
1185: These solutions describe the \textit{non-symmetric} representative planes,  
1186: with respect to the octupole theory. This definition is exact up to the third
1187: order in $\alpha_{i,j}$. Angles $\sigma_1, \sigma_2$ satisfying condition in
1188: Eq.~\ref{condition_non_axial}, i.e., solutions to 
1189: Eq.~\ref{representative_conditions} can be found consistent with higher order
1190: expansions. However,  these equations are very complex and, in practice, we
1191: would have to solve them numerically (for instance, with the Newton-Raphson
1192: algorithm initiated with starting conditions derived from the octupole theory).
1193: Hence, in general, the representation of the energy levels with the help of the
1194: non-symmetric representative planes is much more difficult than in the symmetric
1195: case and is not unique (it has some analogy to the Poincar\'e cross-section).
1196: For instance, we can define the representative plane for higher order expansions
1197: of $\Hsec$. In this paper, we focus on the symmetric representation only.
1198: 
1199: %_______________________________________________________________________________
1200: \subsection{Energy levels for three-planet secular model}
1201: %_______________________________________________________________________________
1202: In this section, to show some applications of the secular theory,  we
1203: investigate qualitative dynamics of a few three-planet extrasolar systems. To
1204: characterize  these systems, we calculated  energy levels in the
1205: \textit{symmetric}  representative plane. We also try to find equilibria in the
1206: secular model of each examined system.
1207: %
1208: \begin{figure*}
1209:  \centerline{
1210:  \vbox{
1211:     \hbox{\includegraphics [width=45mm]{fig3a.eps}\hskip-3mm
1212:           \includegraphics [width=45mm]{fig3b.eps}\hskip-3mm
1213:           \includegraphics [width=45mm]{fig3c.eps}\hskip-3mm
1214:           \includegraphics [width=45mm]{fig3d.eps}}
1215:     \hbox{\includegraphics [width=45mm]{fig3e.eps}\hskip-3mm
1216:           \includegraphics [width=45mm]{fig3f.eps}\hskip-3mm
1217:           \includegraphics [width=45mm]{fig3g.eps}\hskip-3mm
1218:           \includegraphics [width=45mm]{fig3h.eps}}
1219:     \hbox{\includegraphics [width=45mm]{fig3i.eps}\hskip-3mm
1220:           \includegraphics [width=45mm]{fig3j.eps}\hskip-3mm
1221:           \includegraphics [width=45mm]{fig3k.eps}\hskip-3mm
1222:           \includegraphics [width=45mm]{fig3l.eps}}
1223: }}	  
1224:  \caption{
1225:  The secular energy levels on the \textit{\em symmetric  representative plane}
1226: for selected three-planet systems. Map coordinates are  $(x\equiv e_1
1227: \cos{\sigma_1}$, $y \equiv e_2 \cos{\sigma_2})$,  where the secular angles
1228: $\sigma_1,\sigma_2$ are $0$ or $\pi$.   Black thin lines are for energy levels
1229: obtained with different methods: panels in the top row are  for the octupole
1230: theory, panels in the middle row are for the 24th-order expansion derived in
1231: this paper, and  panels in the bottom row are for the semi-analytical averaging.
1232: In gray areas, $e_3<0$, hence the motions are not permitted The light-gray areas
1233: are for the regions of collisions between the inner and the middle planet.  In
1234: orange regions, the secular expansion diverges. The thick, red straight lines
1235: mark the anti-collision lines between  the planets and indicate the true border
1236: of the convergence of the secular expansion. Orbital parameters are taken from
1237: Jean Schneider Encyclopedia, and are given in terms of tuples 
1238:  $\mathbf{p}_i \equiv (m_0 [\mbox{M}_{\odot}],m_1 [\mbox{m}_{\idm{J}}],
1239:  m_2 [\mbox{m}_{\idm{J}}],m_3 [\mbox{m}_{\idm{J}}],
1240:  a_1 [\mbox{AU}],a_2 [\mbox{AU}], a_3 [\mbox{AU}], e_1, e_2, e_3)$ as follows:
1241:   $\mathbf{p}_{\idm{ups And}} = 
1242:  (1.27,0.69,1.98,3.95,0.059,0.83,2.51,0.029,0.254,0.242)$,
1243:  $\mathbf{p}_{\idm{HD~74156}} =
1244:  (1.24,1.88,0.396,8.03,0.294,1.01,3.85,0.64,0.25,0.43)$,
1245:  $\mathbf{p}_{\idm{Gl 581}} = 
1246:  (0.31,0.0492,0.0158,0.0243,0.041,0.073,0.25,0.02,0.16,0.2)$,
1247:  $\mathbf{p}_{\idm{HD~69830}} = 
1248:  (0.86,0.033,0.038,0.058,0.0785,0.186,0.63,0.1,0.13,0.07)$. 
1249:  Each panel is labeled with $AMD$ (expressed 
1250:  in standard units) calculated for the nominal configuration.
1251:  The secular energy levels for the 
1252:  nominal systems are marked with green curves.
1253:  }
1254:  \label{systems}
1255:  \end{figure*}
1256: The results are illustrated in Figure \ref{systems}. We selected four
1257: three-planet configurations. Their orbital elements are taken from the
1258: Extrasolar Planets Encyclopedia of Jean Schneider, following the most recent
1259: determinations of the orbital solutions. Each column in Fig.~\ref{systems} is
1260: for one particular system, i.e., for $\upsilon$~And \citep{Butler2006a},
1261: HD~74156~\citep{Bean2008},  Gliese~581~\citep{Lovis2006a} and 
1262: HD~69830~\citep{Udry2007b}, respectively. Different approximations to the
1263: secular theory are illustrated in rows. The top row panels are for the energy
1264: levels calculated with the octupole theory,  panels  in the middle row are 
1265: derived from the expansion of $\Hsec$ of the $24$-order, and  panels in the
1266: bottom row illustrate the energy levels computed with the numerical algorithm.
1267: The later case may be regarded as the exact solution to the problem thanks to 
1268: adaptive, high order Gauss-Legendre quadratures which we used to compute the double
1269: integral over $\Hpert$. In each case, we fixed $AMD$ consistent with the nominal
1270: parameters of the examined systems.
1271: 
1272: %_______________________________________________________________________________
1273: \subsubsection{$\upsilon$~Andromedae}
1274: %_______________________________________________________________________________
1275: First, we are looking at the exact (numerical) phase plots. In the case of
1276: $\upsilon$~And, we found two types of equilibria. The first one, marked with I,
1277: is the global minimum of the secular Hamiltonian. Hence, according to the
1278: Lyapunov theorem, this equilibrium is  stable (because the Hamiltonian can be
1279: regarded as the positive definite  Lyapunov function).  The equilibrium marked
1280: with~II is a saddle point, and is stable in the linear approximation. It can be
1281: verified by solving the eigenproblem of the linearized equations of motion in
1282: the neighborhood of the equilibrium. Now, we can compare the outcomes of the
1283: analytic theories with the  results of exact, numerical algorithm.  Apparently,
1284: the high-order analytical theory is fully compatible with the numerical theory.
1285: The largest deviations between the secular energies are of the order of
1286: $10^{-9}$. This accuracy is preserved  even for eccentricities $e_1$ close to
1287: $1$. On contrary,  the octupole theory provides only a crude representation of
1288: the phase space.  The phase plot constructed with the help of this theory also
1289: reveals an equilibrium of type I, however, at place of equilibrium II, 
1290: qualitatively different energy levels appear (see the top-left panel  in
1291: Fig.~\ref{systems} with three ``false'' equilibria).
1292: 
1293: To locate the ``real'' system in the energy plot, we mark the level of the
1294: secular energy computed for the nominal parameters of the system with green,
1295: thick curve. It provides only a crude imagination where the system is located;
1296: one should be aware that we are looking at the representative plane (hence
1297: $\sigma_{1,2}$ are fixed at specific values), and we do not take into account 
1298: the parameter errors. Still, the plot tell us that while variability of $e_2$ is
1299: limited, $e_1$ may be varied in all permitted range of eccentricity.
1300: 
1301: %_______________________________________________________________________________
1302: \subsubsection{HD 74156}
1303: %_______________________________________________________________________________
1304: In the phase space of the HD~74156 system,  we discover only one equilibrium
1305: (labeled with III) in the regime of small eccentricities. It is related to the
1306: global maximum of $\Hsec$, and it means that this solution is Lyapunov stable.
1307: The $AMD$ of the nominal system permit eccentricities to reach large values,
1308: hence they enter the regions in which the secular Hamiltonian expansion
1309: diverges (see the explanation in Sect.~2.5). These regions are marked in orange color.  The region of permitted
1310: motions is also bordered by two collision lines of orbits, defined implicitly
1311: through $a_2(1 - e_2) = a_1 ( 1 \pm e_1 )$ and they are marked with gray, thick
1312: lines. In this case the view of the phase plot varies with the order of
1313: expansion (or the applied algorithm). The high-order analytic theory
1314: reconstructs the phase plot for $e_2 \sim [0,0.6)$ (white zone) in almost  whole
1315: permitted range, nevertheless, in the region of divergent expansion the phase
1316: plot is wrong. In the case of octupole theory, we obtain only a crude
1317: approximation of the structure of the phase space, and again the theory
1318: introduces  artifacts (two saddle points and an extremum).
1319: 
1320: We also plot the energy levels of the nominal system (in the same manner as we
1321: did for $\upsilon$~And). Its parameters would evolve along this level relatively
1322: distant from the equilibrium close to the origin.
1323: 
1324: %_______________________________________________________________________________
1325: \subsubsection{Gliese 581 and HD 69830}
1326: %_______________________________________________________________________________
1327: The phase space of systems Gliese~581 and HD~69830 are  quite similar. The
1328: region of permitted motions is limited to relatively small eccentricities.  In
1329: both cases, we have only one equilibrium close to the origin that is related to
1330: the global maximum of the secular Hamiltonian, and therefore they are Lyapunov
1331: stable.  For the Gliese~581 planetary system,  the accuracy of the $24$-order
1332: secular expansion is not very good; at the borders of  permitted motion, this
1333: accuracy is at a level of $10^{-3}$ only.  For the third-order theory, this
1334: accuracy is even worse, $\sim 10^{-2}$. In the case of HD 69830, the relative
1335: accuracy of the high-order expansion is not worse than $2.5 \times 10^{-9}$.
1336: 
1337: %_______________________________________________________________________________
1338: \subsection{Secular dynamics of HD~37124}
1339: %_______________________________________________________________________________
1340: As a particular system to study, we choose  the three-planet system of
1341: HD~37124. It has been discovered by \cite{Vogt2005}. Remarkably,  the most
1342: recent best fit solutions  to the observations are consistent with configurations involving
1343: sub-Jupiter companions in orbits with moderate eccentricities. The eccentricity
1344: of the outermost companion is not well constrained, nevertheless extensive
1345: dynamical analysis of the RV data in \citep{Gozdziewski2007} make it possible
1346: to locate this planet in a region between 8:3 and 11:4~MMRs with the middle 
1347: companion. In that case, the orbital parameters can be regarded as well fitting
1348: the assumptions of the secular theory.
1349: 
1350:  \begin{figure*}
1351:  \centerline{
1352:  \vbox{
1353:     \hbox{\includegraphics [width=54mm]{fig4a.eps}\hskip-2mm
1354:           \includegraphics [width=54mm]{fig4b.eps}\hskip-2mm
1355:           \includegraphics [width=54mm]{fig4c.eps}}
1356:     \hbox{\includegraphics [width=54mm]{fig4d.eps}\hskip-2mm
1357:           \includegraphics [width=54mm]{fig4e.eps}\hskip-2mm
1358:           \includegraphics [width=54mm]{fig4f.eps}}
1359:     \hbox{\includegraphics [width=54mm]{fig4g.eps}\hskip-2mm
1360:           \includegraphics [width=54mm]{fig4h.eps}\hskip-2mm
1361:           \includegraphics [width=54mm]{fig4i.eps}}
1362:          }}
1363:  \caption{
1364:  The secular energy levels on the {\em symmetric  representative plane} for
1365: three-planet system~HD37124 (see the text for more details). The  map
1366: coordinates are $(x \equiv  e_1 \cos{\sigma_1})$ and $(y = e_2
1367: \cos{\sigma_2})$,  where the secular angles are fixed at $0$ (the top
1368: half-plane/the right half-plane) or $\pi$ (the bottom half-plane/the left
1369: half-plane). The boundary between white and grey regions is for $e_3=0$ (i.e.,
1370: the motion is permitted only in the white areas).  Black solid lines are for 
1371: the secular energy  levels obtained with the help of the octupole theory (panels
1372: in the top row),   by the expansion of the $24$-th order (panels in the middle
1373: row)  and  with the semi-numerical averaging (the bottom row of panels). 
1374: Orbital  parameters are taken from the discovery paper (Vogt et al., 2005)
1375: (panels in the left column marked with 1), and from Go\'zdziewski et al. (2007)
1376: (panels in the middle and in the right column, marked with~2 and~3,
1377: respectively).  These parameters, in terms of tuples
1378:  $\mathbf{p}_i   \equiv (m_0 [\mbox{M}_{\odot}],m_1 [\mbox{m}_{\idm{J}}],
1379:  m_2 [\mbox{m}_{\idm{J}}], m_3 [\mbox{m}_{\idm{J}}],
1380:  a_1 [\mbox{AU}], a_2 [\mbox{AU}],
1381:  a_3 [\mbox{AU}], e_1, e_2, e_3)$ are the following: 
1382:  $\mathbf{p}_1 = (0.91,0.61,0.6,0.683,0.53,1.64,3.19,0.055,0.14,0.2)$,
1383:  $\mathbf{p}_2 = (0.78,0.624,0.606,0.581,0.519,1.632,3.212,0.037,0.003,0.048)$,
1384:  $\mathbf{p}_3 = (0.78,0.650,0.584,0.567,0.519,1.668,2.740,0.091,0.040,0.132)$.
1385:  Stationary solutions are labeled with I, II, and~III. Green curves
1386:  are for the secular energy level of the respective nominal initial condition.
1387: }
1388: \label{hd37124}
1389: \end{figure*}
1390:  
1391: Figure~\ref{hd37124} is for the energy levels in the  \textit{symmetric}
1392: representative plane computed and calculated for three slightly different
1393: orbital configurations related to possible orbital best-fits (panels in each 
1394: column are for one orbital fit). Osculating elements of these configurations are
1395: quoted in the caption to this figure. The first, kinematic solution to the
1396: three-Keplerian model of the RV, is taken from the original discovery paper
1397: \citep{Vogt2005}. Other two  best-fits are from \cite{Gozdziewski2007}.  The
1398: energy levels are computed for fixed $AMD$  calculated for each particular
1399: initial condition.
1400: 
1401: In the top row of Fig.~\ref{hd37124} we show energy levels calculated from the
1402: octupole theory, plots in middle row are derived from the expansion of $\Hsec$
1403: of the $24$-order, and the bottom row illustrates the results derived from the
1404: numerical algorithm. In all cases, the analytical high-order theory is in
1405: excellent agreement with the  numerical,  exact theory.  We checked that the
1406: magnitude of largest, relative deviations between the analytic and numerical
1407: results are of the order of $2.5 \times 10^{-6}$. On the other hand, the
1408: octupole theory gives  relatively precise insight into the structure of the
1409: phase space. All qualitative  features  of the energy plane are reproduced 
1410: quite well.
1411: 
1412: The HD~37124 seems to be the most interesting example of the secular dynamics in
1413: the real system found in this paper.  The energy planes reveal unusual dynamical
1414: structures related to the equilibria in the secular system.  We know already
1415: that they can appear as extrema as well as saddle points in the representative
1416: plane.  For the same $AMD$, we can have three types of stationary solutions. Two
1417: of them are characterized by extrema of $\Hsec$ in the four-dimensional phase
1418: space: the first one is the maximum which  appears in the quarter with
1419: $\sigma_{1,2}=0$, and there is  the global minimum of $\Hsec$ in the quarter
1420: with $\sigma_{1}=0$ and $\sigma_{2}=\pi$. We also found  saddle points of
1421: $\Hsec$ in the quarter with  $\sigma_{1}=\pi$, $\sigma_2=0,\pi$. Stationary
1422: points  marked with I and II in Fig.~\ref{hd37124} are related to extrema of
1423: $\Hsec$, hence they are stable. By examining the eigenvalues of the linearized
1424: equations of motions in the neighborhood of the saddle point (equilibrium~III),
1425: we checked out that  it is linearly stable. It can be localized in the
1426: half-plane of  $\sigma_2=0$ or $\sigma_2=\pi$, depending on selected orbital
1427: parameters. 
1428: 
1429: All these solutions appear in the range of moderate eccentricities, and in fact
1430: can be located close to the actual positions of the best fit solutions. The
1431: structure of the energy plane is also robust with respect to small changes of
1432: the orbital parameters. In each panel, similarly to the previous systems, we 
1433: mark the energy level of the respective nominal configuration  with the  green
1434: thick curve. Curiously, depending on the chosen fit, the nominal system can
1435: evolve in the quarter of the representative plane characterized by librations of
1436: $\sigma_{1,2}$ around $0$ (the top-right quarter), or librations of $\sigma_{1}$
1437: around $0$ and $\sigma_{2}$ around $\pi$ (the bottom-right quarter), as well as
1438: $\sigma_{1}$ around $\pi$ while $\sigma_{2}$  can be librating around $0$  or
1439: $\pi$ (the top-left or the bottom-left quarter). It means, that the apses of 
1440: two innermost companions can be all aligned with the apsidal line of the
1441: outermost planet,  they can be also anti-aligned or the apsidal directions can
1442: be mixed.
1443: 
1444: The presence of these stationary solutions can be interpreted  in terms of the
1445: two-planet theory. We recall that for the case of two planets, mode~I (with
1446: apsides aligned)  corresponds to the maximum of the secular energy, while
1447: mode~II (apsides anti-aligned) corresponds to the minimum of $\Hsec$. In the
1448: case of three planets, we add the secular energies of three pairs of interacting
1449: planets. Hence, in a region of the representative plane where  the maxima of
1450: $\Hsec$ of these three planetary pairs can roughly coincide, we can obtain the
1451: maximum of the total energy; by adding $\Hsec$ in the region where the
1452: particular minima are close enough in the parameter space, we can obtain the
1453: global minimum of the energy, and in the case of superimposed minimum and other
1454: maximum we can obtain the saddle of the total energy. Geometrically, the
1455: equilibria can be interpreted as combinations of  the secular modes known from
1456: the theory of nonresonant two-planet system. For instance, the maximum of
1457: $\Hsec$ can be related to triple mode~I (i.e., the neighboring solutions are
1458: characterized with librations $\Delta\varpi$ around~$0$ for all  pairs of
1459: planets), and the saddle point of $\Hsec$ is obtained for a superposition of
1460: mode~I for some pair(s) and of mode~II for the other pair(s).  It is not clear
1461: for us yet, what would mean a combination with  the NSR mode, in the regime of
1462: large eccentricities (i.e., in the region of nonlinear-secular resonance).
1463: Likely, it could be related to sophisticated secular dynamics.
1464: 
1465: 
1466: %_______________________________________________________________________________
1467: \section{Conclusions}
1468: %_______________________________________________________________________________
1469: The number of multi-planet extrasolar systems constantly grows. The Doppler
1470: spectroscopy remains the most effective  detection technique. Unfortunately, the
1471: measurements of RV are in some sense degenerate because due to symmetry of the
1472: Doppler signal, we usually cannot determine the true inclination of planetary
1473: orbits. Other parameters are usually determined with large uncertainties. Hence,
1474: to characterize the dynamics of such systems we cannot rely only on single
1475: initial conditions and effective, qualitative methods  of dynamical analysis are
1476: very desirable. 
1477: 
1478: In this paper we consider the secular theory of a coplanar $N$-planet system
1479: which is far from MMRs and orbital collision zones. In this
1480: case, the high-frequency interactions can be averaged out and we obtain greatly
1481: simplified picture of the long-term behavior of the system. This idea leads to 
1482: the classic Laplace-Lagrange theory and its modern generalizations like the
1483: octupole theory \citep{Ford2000,Lee2003}, high-order expansion of the secular
1484: perturbation \citep{Henrard2005, Gallardo2005} or the semi-numerical  averaging
1485: invented by \cite{Michtchenko2004}. 
1486: 
1487: Our work can be considered as a generalization of the octupole theory for
1488: hierarchical triple systems characterized with large ratio of semi-major axes.
1489: We have shown that in this case the perturbation can be averaged out over mean
1490: longitudes with very basic change of integration variables that makes it
1491: possible to express the integrand function as a polynomial of trigonometric
1492: functions, without any need of relatively complex Fourier expansions.  To the
1493: best of our knowledge, such method has been not applied in the literature.
1494: However, during the final preparation of the manuscript we found a book of
1495: \cite{Valtonen2006}, who use a similar idea to construct the octupole theory of
1496: hierarchical triple stellar systems. 
1497: 
1498: Basically, the secular  Hamiltonian is expressed through the power series with
1499: respect to the ratios of semi-major axes, without an explicit limitation on the
1500: eccentricities. These series can be continued to practically any order. However,
1501: if we apply the averaging algorithm presented in this work, the convergence
1502: region of these series is usually limited. To avoid this problem, in this paper
1503: we also propose  a further improvement of this method and to
1504: generalize the expansion to the spatial problem. Our theory significantly improves the
1505: octupole theory of two-planet systems. We have shown that it can be generalized
1506: to any $N$-planet system fulfilling the assumption of the averaging theorem. The
1507: simple ``trick'' of choosing the integration variables can be applied not only
1508: for purely gravitational point-to-point interactions but also in  other models
1509: in which  the mutual interactions can be expressed in powers of mutual distance
1510: between objects in the system. For instance, now we work on  applying the
1511: averaging algorithm to  relativistic and quadrupole moment perturbations
1512: \citep{Migaszewski2008c}. Its generalization to the 3D problem (in particular,
1513: for two-planet system) is also straightforward. Then it can be applied to the
1514: study of secular dynamics in hierarchical triple-star systems or star--planet
1515: configurations fulfilling  assumptions of the secular theory.
1516: 
1517: In this work, the secular theory is used to investigate stationary solutions in
1518: the three-planet systems that are  relatively frequent in the known sample of
1519: extrasolar planets. We found that the libration modes known in two-planet
1520: configurations can be generalized for the multi-planet model.  Still, our study
1521: of particular systems is quite preliminary and new, yet unknown stationary
1522: solutions are expected to exist in this problem.
1523: 
1524: %_______________________________________________________________________________
1525: \section*{Acknowledgments}
1526: %_______________________________________________________________________________
1527: We thank Tatiana Michtchenko for a detailed review 
1528: and corrections that improved the manuscript.
1529: This work is supported by the Polish Ministry of Sciences and Education, Grant
1530: No. 1P03D-021-29. C.M. is also supported by Nicolaus Copernicus University
1531: Grant No.~408A.
1532: %_______________________________________________________________________________
1533: \bibliographystyle{mn2e}
1534: \bibliography{ms}
1535: \label{lastpage}
1536: \end{document}
1537: 
1538: