0809.5248/ms.tex
1: \documentclass[useAMS,usenatbib,usegraphicx]{mn2e}
2: \usepackage{ifpdf}
3: \usepackage{pdfscreen}
4: %-------------------------------------------------------------------------------
5: % preamble
6: %-------------------------------------------------------------------------------
7: \usepackage{pslatex}
8: \usepackage{amsmath,amsfonts}
9: \newcommand{\Hfull}{\mathcal{H}}
10: \newcommand{\Hsec}{\mathcal{H}_{\idm{sec}}}
11: \newcommand{\Hrel}{\mathcal{H}_{\idm{GR}}}
12: \newcommand{\Hflat}{\mathcal{H}_{\idm{QM}}}
13: \newcommand{\Hbulge}{\mathcal{H}_{\idm{TB}}}
14: \newcommand{\HflatS}{\mathcal{H}_{\idm{spin}}}
15: \newcommand{\HflatT}{\mathcal{H}_{\idm{TB}}}
16: \newcommand{\HsecJ}{\mathcal{H}^J_{\idm{sec}}}
17: \newcommand{\Hpert}{\mathcal{H}_{\idm{pert}}}
18: \newcommand{\Hmut}{\mathcal{H}_{\idm{NG}}}
19: \newcommand{\HpertJ}{\mathcal{H}^J_{\idm{pert}}}
20: \newcommand{\HKepl}{\mathcal{H}_{\idm{kepl}}}
21: \newcommand{\Hint}{\mathcal{H}_{\idm{int}}}
22: \newcommand{\Rsec}{\mathcal{R}_{\idm{sec}}}
23: \newcommand{\Xint}{\mathcal{X}}
24: \newcommand{\R}{\mathcal{R}}
25: \newcommand{\K}{\mathcal{K}}
26: \newcommand{\RP}{${\cal S}$}
27: \newcommand{\RPm}{${\cal S}_{\pi}$}
28: \newcommand{\RPp}{${\cal S}_0$}
29: \newcommand{\TSR}{TSR}
30: \newcommand{\UE}{UE}
31: \def\vec#1{{\mathbf{#1}}}
32: \def\idm#1{{\mbox{\scriptsize #1}}}
33: %\def\corr#1{{\em #1}}
34: \def\corr#1{{ #1}}
35: %
36: \voffset=-0.4in
37: %-------------------------------------------------------------------------------
38: \title[Generalized secular planetary problem]
39: {Secular dynamics of coplanar, non-resonant planetary system
40: under the general relativity and quadrupole moment perturbations}
41: \author[C. Migaszewski and K. Go\'zdziewski]{
42: Cezary Migaszewski$^{1}$\thanks{E-mail: c.migaszewski@astri.uni.torun.pl} and 
43: Krzysztof Go\'zdziewski$^{1}$\footnotemark[1]\thanks{E-mail: k.gozdziewski@astri.uni.torun.pl}\\
44: $^{1}$Toru\'n Centre for Astronomy, Nicolaus Copernicus University, 
45: Gagarin Str. 11, 87-100 Toru\'n, Poland}
46: \begin{document}
47: %-------------------------------------------------------------------------------
48: 
49: \date{Accepted 2008 September 20.  Received 2008 September 19; in original form
50: 2008 April 14}
51: \pagerange{\pageref{firstpage}--\pageref{lastpage}} \pubyear{2008}
52: 
53: \maketitle
54: 
55: \label{firstpage}
56: 
57: %-------------------------------------------------------------------------------
58: \begin{abstract}
59: We construct a secular theory of a coplanar system of $N$-planets not involved
60: in strong mean motion resonances, and which are far from collision zones.
61: Besides the point-to-point Newtonian mutual interactions, we  consider the
62: general relativity corrections to the gravitational potential of the star and
63: the innermost planet, and also a modification of this potential by the
64: quadrupole moment and tidal distortion of the star. We focus on hierarchical
65: planetary systems.  After averaging the model Hamiltonian with a simple
66: algorithm making use of very basic properties of the Keplerian motion, we obtain
67: analytical secular theory of the high order in the semi-major axes ratio. A
68: great precision of the analytic approximation is demonstrated with the numerical
69: integrations of the equations of motion.  A survey regarding model parameters
70: (the masses, semi-major axes, spin rate of the star) reveals a rich and
71: non-trivial dynamics of the secular system. Our study is focused on its
72: equilibria. Such solutions predicted by the classic secular theory,  which
73: correspond to aligned (mode~I) or anti-aligned (mode~II) apsides, may be
74: strongly affected by the gravitational corrections. The so called true secular
75: resonance, which is a new feature of the classic two-planet problem discovered
76: by Michtchenko \& Malhotra (2004), may appear in other, different regions of the
77: phase space of the generalized model. 
78: \corr{
79: We found bifurcations of mode~II which emerge new, yet unknown in the
80: literature, secularly unstable equilibria and a complex structure of the phase
81: space. These equilibria may imply secularly unstable orbital configurations
82: even for initially moderate eccentricities.
83: } 
84: The point mass gravity corrections can affect the long term-stability in the
85: secular time scale,  which may directly depend on the age of the host star
86: through its spin rate. We also analyze the secular dynamics of the
87: $\upsilon$~Andromede system in the realm of the generalized model. Also in this
88: case of the three-planet system, new secular equilibria may appear.
89: \end{abstract}
90: %-------------------------------------------------------------------------------
91: 
92: \begin{keywords}
93: celestial mechanics -- secular dynamics -- relativistic effects -- 
94:  quadrupole moment -- analytical methods -- stationary
95: solutions --  extrasolar planetary systems
96: \end{keywords}
97: 
98: %-------------------------------------------------------------------------------
99: \section{Introduction}
100: %-------------------------------------------------------------------------------
101: The currently known sample of extrasolar planets\footnote{See
102: www://exoplanet.eu, http://exoplanets.org, http://exoplanets.eu} comprises of
103: many multiple-planet configurations. Their architectures are diverse and,
104: usually, very different from the Solar system configuration. Some of them
105: consist of so called hot-Jupiters or hot-Neptunes, with  semi-major axes $\sim
106: 0.05$~au and the orbital periods of a few days.  The multiple-planet systems may
107: also contain companions in relatively distant orbits. Such systems are 
108: non-resonant (or hierarchical). 
109: 
110: The discovery of multi-planet systems and their unexpected orbital diversity
111: initiated interest in their secular evolution. Generally, the analysis of
112: secular interactions, including tidal effects, general relativity (GR), and
113: quadrupole moment (QM) corrections to the Newtonian gravitation (NG) of point
114: masses, follows the theory of compact multiple stellar systems
115: \citep{Mardling2002,Nagasawa2005}.  The secular evolution  of planetary systems
116: has been already intensively studied in this context by many authors. For
117: instance, \cite{Adams2006a} consider the GR corrections in the sample of
118: detected extrasolar systems. They have shown that the GR interactions can be particularly 
119: important for the long-term dynamics of  short-period planets. Also
120: \cite{Adams2006b,Mardling2007} study the evolution of such objects taking into
121: account the tidal circularization of their orbits, and secular excitation of the
122: eccentricity of the innermost planet by more distant companions. These secular
123: effects imply constraints on the orbital elements of the innermost orbits,
124: including yet undetected Earth-like planets, which are expected to be found in
125: such systems. The  quadrupole moment  and relativistic effects may detectably
126: affect the light-curves of stars hosting transiting planets in a few years
127: time-scale  \citep{MiraldaEscude2002,Agol2005,Winn2005}.  In  general, the
128: interplay of apparently subtle perturbations with the  point mass Newtonian
129: gravity depends on many physical and orbital parameters, and it may lead to
130: non-trivial and interesting dynamical phenomena. \corr{Still, 
131: the problem is of particular
132: importance for studies of the long-term stability of the Solar system
133: \citep[see very recent works by][]{Benitez2008,Laskar2008}.}
134: 
135: Because a fully general study of such effects would be a very difficult task, we
136: introduce some simplification to the planetary model. We skip tidal effects
137: caused by dissipative interactions of extended star and planetary bodies, in
138: particular  the tidal friction (TF) damping the eccentricity and the tidal
139: distortion (TD) that modifies planetary figures,  and could influence their
140: gravitational interaction with the parent star. The tidal effects are typically
141: much smaller than the leading Newtonian, relativistic and quadrupole moment
142: contributions \citep{Mardling2007}. Also their time-scale is usually much longer
143: than of the most  significant conservative effects.    In this work, we focus on
144: the secular dynamics of planetary systems over an intermediate time scale
145: relevant to the  conservative perturbations.  Our goal is to obtain a
146: qualitative picture of the secular system that may be useful as the first order
147: approximation to more general model of the long-term planetary dynamics. 
148: 
149: Moreover, we show in this paper that even if we skip the non-conservative  tidal
150: effects then the GR and/or QM {\em corrections} to the 
151: point mass Newtonian gravity can be
152: much more significant for the secular dynamics than the mutual NG interactions
153: alone. These effects may change qualitatively  the view of the dynamics of
154: planetary systems  as predicted by the Laplace-Lagrange theory
155: \citep{Murray2000} or its recent versions
156: \citep[e.g.,][]{Lee2003,Michtchenko2004,Michtchenko2006,Libert2005,Veras2007,Migaszewski2008a}. 
157: 
158: The model without dissipative effects may be investigated  with the help of
159: conservative Hamiltonian theory. Assuming that planetary orbits are well
160: separated and the system is far from mean motion resonances and collision zones,
161: we can apply the averaging proposition \citep{Arnold1993} to derive the long
162: term evolution of their orbital elements. This approach can be classified among
163: secular theories having a long history in the context of the Solar system
164: \citep{Brouwer1961,Murray2000}.  
165: 
166: The present work is a step towards a generalization of the secular model of
167: coplanar 2-planet system by \cite{Michtchenko2004} and its analytical version
168: applied to $N$-planet system in \citep{Migaszewski2008a}.  To explore the phase
169: space globally without restrictions on eccentricities, we simplify the equations
170: of motion through the averaging of perturbations to the Keplerian motion, with
171: the help of semi-numerical method proposed by \cite{Michtchenko2004}. To obtain
172: analytical results, we also use  a very simple averaging algorithm that can be
173: applied to perturbations dependent on the mutual distance of interacting bodies
174: \citep{Migaszewski2008a}. These works demonstrate that the secular evolution can
175: be precisely described in wide ranges of the orbital parameters, including
176: eccentricities up to 0.8--0.9, for well separated (hierarchical)  orbits with
177: small ratio of the semi-major axes, $\alpha\sim 0.1$.  Here, the secular NG
178: theory is very helpful to derive 
179: the more general model including the GR and QM  effects. Basically, without the
180: averaging, the only possibility of investigating the long-term secular dynamics
181: relies on numerical solutions of the equations of motion \citep{Mardling2002}.
182: However, due to extremely different time scales which are related as days (the
183: orbital periods of inner planets) to $10^3$--$10^6$~years of the secular orbital
184: evolution, the numerical integrations are of very limited use when we want to
185: investigate large volumes of initial conditions rather than isolated orbits. In
186: such a case,  the analytical theory can be very helpful to get a deep insight
187: into the secular dynamics. Moreover, when necessary, particular solutions can be studied
188: in detail with the help of the direct numerical integrations.
189: 
190: The plan of this paper is the following. In Sect.~2 we formulate the generalized
191: model of a coplanar, $N$-planet system, accounting for the relativistic and
192: quadrupole moment corrections to the NG-perturbed motion of the innermost
193: planet. To make the paper self-consistent, we average out the perturbations to
194: the Keplerian motion  with the help of the averaging algorithm described in
195: \citep{Migaszewski2008a}. In Sect.~3 we apply the analytical and numerical tools
196: to study the influence of the GR and QM corrections on the secular evolution. 
197: In particular, we recall the concept of the so called representative  plane of
198: initial conditions \citep{Michtchenko2004}. We focus on the  search for
199: stationary solutions in the averaged and reduced systems (periodic orbits in the
200: full systems), and we investigate their stability and bifurcations with the help
201: of phase diagrams.  In this section we also derive interesting conclusions on
202: the stability of the unaveraged systems. In Sect.~4, we study the phase space of
203: two-planet systems in wide ranges of parameters governing their orbital
204: configurations (masses, semi-major axes ratios, eccentricities) and physical
205: parameters (flattening of the star). To illustrate the application of the
206: secular theory to multi-planet configurations, we consider the three-planet
207: $\upsilon$~Andr system, and we investigate the QM (stellar rotation) influence
208: on its secular orbital evolution.
209: %
210: %-------------------------------------------------------------------------------
211: \section{Generalized model of $N$-planet system}
212: %-------------------------------------------------------------------------------
213: %
214: The dynamics of the planetary system can be modeled by the Hamiltonian function
215: written with respect to canonical Poincar\'e  variables \citep[see,
216: e.g.,][]{Poincare1897,Laskar1995}, and    expressed by a sum of two terms,
217: $
218: \Hfull = \HKepl + \Hpert,
219: $
220: where
221: \begin{equation}
222: \HKepl = \sum_{i=1}^{N} {\bigg( \frac{\mathbf{p}_i^2}{2 \beta_i} 
223: - \frac{\mu_i \beta_i}{r_i} \bigg)}
224: \label{HKepl}
225: \end{equation}
226: stands for  integrable part comprising of the direct sum of the relative,
227: Keplerian motions of $N$ planets and the host star. Here, the dominant point
228: mass of the star is ${m_0}$, and $m_i \ll m_0$, $i=1,\ldots,N$ are the point
229: masses of the $N$-planets.  For each planet--star pair we define the mass
230: parameter ${\mu_i=k^2~(m_0+m_i)}$ where $k$ is the Gauss gravitational constant,
231: and ${\beta_i=(1/m_i+1/m_0)^{-1}}$ are the so called reduced masses. We consider
232: the perturbing Hamiltonian  $\Hpert$ as a sum of three terms, 
233: \begin{equation}
234: \Hpert = \Hmut + \Hrel + \Hflat,
235: \label{Hpert}
236: \end{equation}
237: where $\Hmut$ is for the mutual point-mass interactions  between planets,
238: $\Hrel$ is for the general (post-Newtonian) relativity corrections to the
239: Newtonian gravity, and $\Hflat$ takes into account the dynamical flattening and
240: tidal distortion of the parent star. 
241: 
242: We consider the secular effects of $\Hmut$, which can be expressed as follows:
243: \begin{equation}
244: \Hmut = \sum_{i=1}^{N-1} \sum_{j>i}^{N} {\bigg(
245:  - \underbrace{\frac{k^2 m_i m_j}{\Delta_{i,j}}}_{\textrm{\small direct part}} +
246: \underbrace{\frac{\mathbf{p}_i \cdot \mathbf{p}_j}{m_0}}_{\textrm{\small indirect
247: part}}\bigg)},
248: \label{Hmut}
249: \end{equation}
250: where  ${\mathbf{r}_i}$ are for the position vectors of planets relative to the
251: star,  ${\mathbf{p}_i}$ are for their conjugate momenta relative to the {\em
252: barycenter} of the whole $(N+1)$-body system, 
253: ${\Delta_{i,j}=\|\mathbf{r}_i-\mathbf{r}_j\|}$ denote the relative distance
254: between planets $i$  and $j$.
255: 
256: For  hierarchical planetary  systems (weakly interacting binaries), the secular
257: time scale of Newtonian point-mass interactions may be as short as 
258: $10^{3}$~years, up to Myrs. In general, these interactions cause slow
259: circulation of the apsidal lines. The GR correction to the NG potential of the
260: star--planet system also leads to the circulation of pericenters.
261: Moreover, the time scale of this effect may be comparable to that one forced by
262: mutual Newtonian interactions between planets. In  such a situation, one should
263: necessarily include the relativistic corrections to the model of motion.
264: %
265: %-------------------------------------------------------------------------------
266: \subsection{General relativity corrections}
267: %-------------------------------------------------------------------------------
268: %
269: The GR correction will be applied only to the innermost planet, so we skip the
270: direct GR perturbations on the motion of the more distant companions, as well as
271: the mutual GR interactions caused by planetary masses. Usually, the GR 
272: corrections to the Newtonian potential are expressed in terms of the PPN
273: formalism \cite[e.g.,][]{Kidder1995}. Alternatively, we found a very clear paper
274: of \cite{Richardson1988} {conveniently providing explicit Hamiltonian of the two
275: body problem with the GR term}. Following these authors, $\Hrel' \equiv
276: \Hrel/\beta$ (i.e., $\Hrel$ rescaled by the  reduced mass) may be written as
277: follows:
278: \begin{equation}
279: \Hrel' = \gamma_1 \vec{P}^4 + \gamma_2 \frac{\vec{P}^2}{r} + \gamma_3
280: \frac{\left(\vec{r} \cdot \vec{P}\right)^2}{r^3} + \gamma_4 \frac{1}{r^2},
281: \label{Hrel}
282: \end{equation}
283: where $\gamma_1,\gamma_2,\gamma_3,\gamma_4$ are coefficients defined 
284: through
285: \[
286: \gamma_1 = - \frac{\left(1 - 3 \nu\right)}{8 c^2}, \quad
287:    \gamma_2 = - \frac{\mu \left(3 + \nu\right)}{2 c^2},\quad
288: \gamma_3 = \frac{\mu^2}{2 c^2}, \quad
289:   \gamma_4 = - \frac{\mu \nu}{2 c^2},
290: \]
291: and  $c$ is the velocity of light in a vacuum, $\mu = k^2 (m_0 + m_1)$,  $\nu
292: \equiv m_0 m_1 / (m_0 + m_1)^2$, $r$ is the astrocentric distance, and
293: $\vec{P}$ is  the astrocentric momentum of the
294: innermost planet (normalized through the reduced mass): 
295: \begin{equation} \vec{P} = \vec{v} + \frac{1}{c^2}
296: \left[ 4\gamma_1 (\vec{v} \cdot \vec{v}) \vec{v} + 
297: \frac{2\gamma_2}{r} \vec{v} + \frac{2\gamma_4}{r^3} (\vec{r} \cdot \vec{v})
298: \vec{r} \right],
299: \end{equation} 
300: where $ \vec{v} \equiv \dot{\vec{r}}$ stands for the  astrocentric velocity of
301: the innermost planet (the relativistic corrections from  other  planetary bodies
302: in the system are skipped). Hence, in the relativistic Hamiltonian, we put 
303: $\vec{P} = \vec{v}$ with the accuracy of $\vec{O}({c^{-2}})$ and then the
304: Hamiltonian is conserved up to the order of $O({c^{-4}})$.
305: %
306: %-------------------------------------------------------------------------------
307: \subsection{Rotational and tidal distortions of the star}
308: %-------------------------------------------------------------------------------
309: %
310: Fast rotating stars are significantly flattened and that in turn may lead to
311: important deviations of the NG potential. In the absence of a close planet
312: and  non-radial pulsations,  the star has the rotational symmetry, regarding its
313: mass density and shape. The gravitational potential of such a body can  be
314: expanded in harmonic series expressed in terms of  the Stokes coefficients. The
315: well known property of this expansion is that the gravitational potential of
316: rotationally symmetric bodies retains only terms related to the so called {\em
317: zonal harmonics}:
318: \begin{equation}
319: \HflatS = \frac{\mu \beta}{r} \sum_{l=2}^{\infty} 
320: {J_{l} \left(\frac{R_0}{r}\right)^{l}} P_l(\sin\phi),
321: \label{Hflat}
322: \end{equation}
323: where $R_0$ stands for the characteristic radius of a sphere encompassing the
324: body, and it can be fixed as the equatorial radius of the star, $P_l(\sin\phi)$
325: is the Legendre polynomial of the $l$-th order, $\phi$ is the astrocentric
326: latitude and $J_l$ are non-dimensional Stokes coefficients, $l>1$. It can be
327: shown that for rotationally symmetric objects, $J_l=0$ for $l$ odd.  Basically,
328: $J_l$ (and the leading term with $J_2$, in particular) can be determined
329: numerically  with the help of the theory of stellar interiors combined with a
330: model of rotation and helio-seismic data \citep{Godier1999,Pijpers1998}.  To
331: calculate these coefficients, one must know the mass density as well as the
332: hydrostatic figure of the star.  There are attempts to estimate $J_2$ for stars
333: hosting planets.  For instance, \cite{Iorio2006} determined quadrupole moment of
334: HD~209458~\citep{Charbonneau2000} as $J_2 \sim 3.5 \times 10^{-5}$, 
335: \corr{however with large error of $\sim 10^{-3}$, making the result not very
336: credible}.  In general, the estimates of $J_{l}$ are very  uncertain even for  the
337: well explored Sun. For this slowly rotating dwarf (with rotational period of
338: $\sim 30$~days), $J_2 \sim 10^{-7}$ with an upper bound of $10^{-6}$. Moreover,
339: in the literature we found approximations of current  $J_2$ of the Sun in the
340: range up to $10^{-5}$. 
341: 
342: Due to indefiniteness of higher order zonal harmonics, we consider the 
343: dynamical effects of the leading $J_2$ term only.  For a coplanar system, in
344: which planets move around the star in its equatorial plane,  the first non-zero
345: term in harmonic series, Eq.~\ref{Hflat}, has the form of:
346: \begin{equation}
347: \label{eq:hrf}
348: \HflatS \approx -\frac{1}{2} \mu \beta J_2 R_0^2 \frac{1}{r^3}.
349: \end{equation}
350: The above formulae may be conveniently expressed 
351: \citep{Mardling2002} in terms of
352: the stellar spin frequency:
353: \begin{equation}
354: \HflatS \approx -\frac{1}{6} \beta R_0^5 \left(1 + m_1/m_0\right) k_L \Omega^2
355: \frac{1}{r^3},
356: \end{equation}
357: where $k_L$ is the tidal Love number \citep{Murray2000} 
358: which can be related to $J_2$ through:
359: \begin{equation}
360: k_L = 3 \frac{J_2}{q}, \quad q = \frac{\Omega^2 R_0^3}{k^2 m_0},
361: \end{equation}
362: $q$ is the ratio of centrifugal acceleration and gravitation at the
363: surface of the star, and $\Omega$ is the spin rate.
364: Hence, we have:
365: \begin{equation}
366: J_2 = \frac{1}{3} k_L \frac{R_0^3 \Omega^2}{k^2 m_0}.
367: \end{equation}
368: In this work we adopt the standard value of $k_L=0.02$ \citep{Nagasawa2005}.
369: Obviously, the dynamical flattening is more significant for fast rotating stars.
370: Before the Zero Age Main Sequence (ZAMS) stage, the rotational periods may be as
371: low as a few days, down to 1~day for young ($\sim 100$~Myr) Sun-like stars. The
372: rotational periods of $\sim 40$~days are typical for 8~Gyrs old, evolved
373: objects.  According to the scaling rule of  $J_2 \sim \Omega^2$, the zonal
374: harmonics of young objects may be as large as $10^{-4}$.   Because the
375: rotational period may change by two orders of magnitude during the life-time of
376: the star, also the quadrupole moment may change by a few orders of magnitude. 
377: Hence, flattening caused by fast rotation not only can affect significantly the
378: orbital acceleration which can compete with the GR and  mutual NG contributions
379: but it may also introduce a dependence of the system dynamics on the age of the
380: parent star \citep[][see also Sect.~5 and Sect.~6]{Nagasawa2005}.
381: 
382: We also consider a contribution of the tidal bulge (TB) caused by  the presence
383: of point-mass innermost planet, assuming that only the extended star is
384: distorted due to the tidal interactions. Then the Hamiltonian is corrected by the
385: following term \citep{Mardling2002}:
386: \begin{equation}
387: \label{eq:htb}
388: \HflatT = -\beta R_0^5 k^2 m_1 \left(1 + m_1/m_0\right) k_L \frac{1}{r^6}. 
389: \end{equation}
390: Both gravitational corrections are then $\Hflat = \HflatS + \HflatT$.
391: 
392: We account for direct dynamical effects related to $\Hflat$ only in the motion
393: of the innermost planet although these perturbations can be also  quite easily
394: included for the remaining star-planet pairs.  In the realm of the secular
395: theory of non-resonant, hierarchical systems, we assume that other companions
396: are much more distant from the star. For the semi-major axes ratio $\sim 0.1$,
397: the GR+QM perturbations acting on such outer companions are by orders of
398: magnitude smaller than for the innermost planet. However, the dynamics of the
399: innermost body influences  indirectly the secular dynamics of the whole system.
400: This will be demonstrated in this work in the cases of two- and three-planet
401: configurations. We also underline that with the simplified  model of
402: interactions, we can study the dynamics of quite compact star--planets 
403: configurations  because they can be parameterized, in general, by 
404: individual planetary
405: masses, semi-major axes, and physical parameters of the star.
406: %
407: %-------------------------------------------------------------------------------
408: \subsection{The secular model of non-resonant planetary system}
409: %-------------------------------------------------------------------------------
410: %
411: To apply the canonical perturbation theory, we first transform $\Hfull$ to the
412: following form:
413: \begin{equation}
414:  \Hfull(\vec{I},\vec{\phi}) = \HKepl(\vec{I}) + \Hpert(\vec{I},\vec{\phi}),
415: \label{poincare} 
416: \end{equation}
417: where $(\vec{I},\vec{\phi})$ stand for the action-angle variables, and
418: $\Hpert(\vec{I},\vec{\phi}) \sim \epsilon \HKepl(\vec{I})$,  where $\epsilon \ll
419: 1$ is a small parameter. In this work, we apply the well known 
420: approach to analyze the equations of motion induced by Hamiltonian,
421: Eq.~\ref{poincare}, that relies on the averaging proposition \citep[see,
422: e.g.,][]{Arnold1993}.  By averaging the perturbations with respect to the fast
423: angles ({\em the mean longitudes} or {\em the mean anomalies}) over their
424: periods,  we obtain the secular Hamiltonian which does not depend on these fast
425: angles. Simultaneously,  the  conjugate momenta to the fast angles become
426: integrals of the secular problem.  In the planetary system with a dominant
427: stellar mass, the orbits (apsidal lines) also slowly rotate due to mutual
428: interactions, hence the longitudes of periastron and the longitudes of node
429: become slow angles. Assuming that no strong  mean motion resonances are present,
430: and the system is far enough from collisions, the averaging makes it possible to
431: reduce the number of the degrees of freedom, and to obtain qualitative
432: information on the long-term changes of the slowly varying orbital elements
433: (i.e., on the slow angles and their conjugate momenta).
434: 
435: The transformation of the Hamiltonian to the required form (Eq.~\ref{poincare})
436: may be accomplished   by expressing this Hamiltonian with respect to the
437: (modified) Delaunay canonical elements. These variables can be related to the
438: Keplerian canonical elements \citep{Murray2000}. Actually, we use the following
439: set of canonical action-angle variables which can be obtained after an
440: appropriate canonical transformation of the Delaunay elements:
441: \begin{eqnarray}
442: \label{dvars}
443: {l_i \equiv \mathcal{M}_i}, & \quad {L_i=\beta_i~\sqrt{\mu_i~a_i}},\nonumber\\
444: {g_i \equiv \varpi_i}, & \quad  {G_i=L_i~\sqrt{1-e_i^2}},\\
445: {h_i \equiv \Omega_i}, & \quad{H_i=G_i~(\cos~I_i}-1),\nonumber
446: \end{eqnarray}
447: where $\mathcal{M}_i$ are the mean anomalies,  $a_i$ stand for canonical
448: semi-major axes,  $e_i$ are the eccentricities,   $I_i$ denote inclinations,
449: $\varpi_i$ are the longitudes of pericenter, and $\Omega_i$ denote the
450: longitudes of ascending node. The choice of $\varpi_i$ instead of $\omega_i$ is
451: important for further applications, because $\Omega_i$ are undefined (and
452: irrelevant) for the dynamics of the coplanar system. We note that the
453: geometrical, canonical elements ($a_i$, $e_i$, $I_i$, $\varpi_i$, $\Omega_i$)
454: may be derived through the formal transformation between classic (astro-centric)
455: Keplerian elements and the relative  Cartesian coordinates
456: \citep[e.g.,][]{FerrazMello2006a,Morbidelli2003}, with appropriate rescaling of the astrocentric
457: velocities. In the settings adopted here, the Cartesian  coordinates are
458: understood as Poincar\'e coordinates, i.e., {\em astrocentric} positions of
459: planets, and canonical momenta taken relative to the {\em barycenter} of the
460: system. 
461: 
462: The $N$-planet Hamiltonian expressed in terms of the modified  Delaunay variables
463: (\ref{dvars}) has the form of:
464: \[
465: \Hfull = -\sum_{i=1}^N \frac{\mu_i^2 \beta^3_i}{2 L_i^2} +
466: \Hpert\underbrace{(L_i,l_i,G_i,g_i,H_i,h_i)}_{i=1,\ldots,N}.
467: \]
468: In this Hamiltonian, $l_i$ play the role of the fast angles. In the absence of
469: strong mean motion resonances, these angles can be eliminated by the following
470: averaging formulae:
471: \begin{eqnarray}
472: \label{secular}
473: && \Hsec=\frac{1}{(2 \pi)^N}\underbrace{\int_0^{2 \pi} \ldots 
474: \int_0^{2\pi}}_{i=1,\ldots,N}{\Hmut \, d\mathcal{M}_1 \ldots d\mathcal{M}_N} +\\
475: && \quad + \frac{1}{2\pi} \int_0^{2\pi}{\Hrel \,d\mathcal{M}_1} 
476: + \frac{1}{2\pi} \int_0^{2\pi}{\Hflat \,d\mathcal{M}_1}\nonumber.
477: \end{eqnarray}
478: Hence, we can rewrite the secular Hamiltonian in the symbolic form of:
479: \begin{equation}
480: \Hsec = \left<\Hmut\right> + \left<\Hrel\right> + \left<\Hflat\right>.
481: \label{Hsec}
482: \end{equation}
483: Now, we try to average out each component of this Hamiltonian expressed  with
484: respect to the fast angles (the
485: mean anomalies).
486: %
487: %-------------------------------------------------------------------------------
488: \subsection{Averaging Newtonian point-to-point interactions}
489: %-------------------------------------------------------------------------------
490: %
491: A simple averaging of  the Hamiltonian of the classic planetary
492: $N+1$-body model is described in our previous paper \citep{Migaszewski2008a}. 
493: The algorithm makes use on the very basic properties of the Keplerian motion and
494: it relies on appropriate change of integration variables in Eq.~\ref{secular}.
495: It may be also applied to  perturbations expressed through powers of the
496: relative distance.
497: 
498: To recall the main result, the secular Hamiltonian  of the $N$-planet system can be
499: described as a sum of two-body Hamiltonians evaluated over all pairs of planets:
500: \begin{equation}
501: \left<\Hmut\right> = \sum_{i=1}^{N-1}  \sum_{j>i}^{N}{\left<\Hmut^{(i,j)}\right>}.
502: \label{avHmut}
503: \end{equation}
504: The direct part of the disturbing Hamiltonian reads as follows:
505: \begin{eqnarray}
506: \label{expanssion}
507: & & \left<\Hmut^{(i,j)}\right> = -\frac{k^2 m_i m_j}{a_j} \times  \nonumber \\
508: && \quad \times \left[1 + \sqrt{1-e_j^2} \sum_{l=2}^{\infty}
509: {\left(\frac{\alpha_{i,j}}{1-e_j^2}\right)^l
510: \mathcal{R}^{(i,j)}_l(e_i,e_j,\Delta\varpi_{i,j})}\right].
511: \end{eqnarray}
512: The explicit formulae for functions $\R^{(i,j)}_l(e_i,e_j,\Delta\varpi_{i,j})$ are
513: given in \citep{Migaszewski2008a}. It is well known that the indirect part
514: averages out to a constant and it does not contribute to the secular dynamics
515: \citep{Brouwer1961}. 
516: %
517: %-------------------------------------------------------------------------------
518: \subsection{Averaging the PPN relativistic potential}
519: %-------------------------------------------------------------------------------
520: %
521: Making use of the formulae for the relativistic PPN Hamiltonian by
522: \cite{Richardson1988}, we write down the mean relativistic potential as follows:
523: \begin{equation}
524: \left<\Hrel'\right> = \gamma_1 \left<\vec{v}^4\right> + \gamma_2
525: \left<\frac{\vec{v}^2}{r}\right> + \gamma_3 \left<\frac{\left(\vec{r} \cdot
526: \vec{v}\right)^2}{r^3}\right> + \gamma_4 \left<\frac{1}{r^2}\right>,
527: \label{Hrelsec}
528: \end{equation}
529: with the accuracy of $O(c^{-2})$.
530: We should average out each component of this Hamiltonian over the mean
531: anomalies. It appears to be possible with  the algorithm in
532: \citep{Migaszewski2008a}. To average out the whole PPN Hamiltonian, we must
533: calculate a  few integrals  written in the general form of:
534: \begin{equation}
535: \left<\Xint\right> \equiv \frac{1}{2\pi} \int_0^{2\pi} {\Xint \,
536: d\mathcal{M}}
537: \equiv \frac{1}{2\pi} \int_0^{2\pi} {\left[\Xint \, {\cal J}\right] \, df},
538: \end{equation}
539: where ${\cal J}$ is a scaling function defined with: 
540: \begin{equation}
541: {\cal J} \equiv \frac{d\mathcal{M}}{df} = \frac{\left(1 - e^2\right)^{3/2}}{\left(1 + e
542: \cos{f}\right)^2},
543: \end{equation}
544: and $f$ is the true anomaly of the innermost planet (note that we skip index
545: "1" of that planet). Components  of the
546: mean Hamiltonian can be written explicitly as follows:
547: \begin{eqnarray}
548: \label{v4}
549: &&\left<\vec{v}^4\right> = \frac{n^4 a^4}{\sqrt{1 - e^2}} \left[1 + \frac{3}{2}
550: e^2 + e^4 \mathcal{F}_1\right],\\
551: \label{v2r}
552: &&\left<\frac{\vec{v}^2}{r}\right> = \frac{n^2 a}{\sqrt{1 - e^2}} \left[1 + e^2
553: \mathcal{F}_2\right],\\
554: \label{rvr3}
555: &&\left<\frac{\left(\vec{r} \cdot \vec{v}\right)^2}{r^3}\right> = \frac{n^2 a}{\sqrt{1 - e^2}} e^2
556: \mathcal{F}_2,\\
557: \label{r2}
558: &&\left<\frac{1}{r^2}\right> = \frac{1}{a^2 \sqrt{1 - e^2}}.
559: \end{eqnarray}
560: Functions $\mathcal{F}_1$ and $\mathcal{F}_2$ are given through:
561: \begin{eqnarray}
562: && \mathcal{F}_1 = \frac{1}{2\pi} \int_0^{2\pi} {\frac{\sin^4{f}}{\left(1 + e
563: \cos{f}\right)^2}~df},\\
564: && \mathcal{F}_2 = \frac{1}{2\pi} \int_0^{2\pi} {\frac{\sin^2{f}}{1 + e \cos{f}}~df}.
565: \end{eqnarray}
566: These integrals can be calculated in the closed form:
567: \begin{equation}
568: \mathcal{F}_1 = \frac{3 \left(2 - e^2 - 2\sqrt{1-e^2}\right)}{2 e^4}, \quad
569: \mathcal{F}_2 = \frac{1-\sqrt{1-e^2}}{e^2}.
570: \end{equation}
571: Finally, the secular relativistic potential is the following:
572: \begin{equation}
573: \left<\Hrel'\right> = - \frac{3 \mu^2}{c^2 a^2 \sqrt{1 - e^2}} +
574: \frac{\mu^2 \left(15 - \nu\right)}{8 a^2 c^2}.
575: \label{eq28}
576: \end{equation} 
577: {
578: Because all secular Hamiltonian corrections we account for in this work (see
579: below) do not depend on the  mean anomalies, $L_{1,2}$ are constants of motion. 
580: It also means  that the second term
581: in Eq.~\ref{eq28} reduces to
582: the constant and does not contribute to the long-term dynamics.
583: }
584: Writing down the secular Hamiltonian with respect to  the action-angle
585: variables, we have:
586: \begin{equation}
587: \left<\Hrel'\right> = -\frac{3 \mu^4 \beta^4}{c^2 L^3 G} + \mbox{const}.
588: \label{avHrel}
589: \end{equation}
590: Here, the action variables $(L,G)$ are defined with usual formulae known
591: in the Keplerian motion theory. We recall that in the secular PPN Hamiltonian,
592: only the star-innermost planet interactions are considered.  The  perturbation
593: has been also scaled by the reduced mass. To  properly add the GR star-inner
594: planet  interactions to the Hamiltonian of $N$-planets, we should account for
595: the mass factor. Because the canonical Delaunay elements of the $N$-planet system are
596: defined   through Eq.~\ref{dvars}, hence, to obtain the proper scaling, we should  multiply
597: the two-body Hamiltonian, Eq.~\ref{avHrel}, through $\beta \equiv 1/(1/m_0 +
598: 1/m_1)$.  Finally, we obtain the following secular Hamiltonian related to the GR
599: correction: 
600: \begin{equation} 
601: \left<\Hrel\right> = -\frac{3 \mu^4 \beta^5}{c^2 L^3 G} + \mbox{const}. 
602: \label{avbHrel}
603: \end{equation} 
604: In this Hamiltonian all slow angles are cyclic, 
605: hence $L$ and $G$ would be integrals of motion in the absence of 
606: {mutual planetary interactions}. However, the element $G$ is
607: no longer constant when we consider the NG contributions. 
608: The evolution of canonical angle $\varpi$, reads as follows: 
609: \begin{equation}
610: \dot{\varpi}_{\idm{GR}} = \frac{\partial \left<\Hrel\right>}{\partial G} =
611: \frac{3 \mu^4 \beta^5}{c^2 L^3 G^2} = \frac{3 \mu^{3/2}}{c^2 a^{5/2} \left(1 -
612: e^2\right)}. 
613: \label{orel}
614: \end{equation} 
615: This is the well known formulae describing the  {\em
616: relativistic advance of pericenter}. We derive it here to keep the paper
617: self-consistent.  In the above formulae,  we skip  the symbol
618: $\left<\ldots\right>$ of the mean, which basically should encompass the
619: canonical angle $\varpi$  and the conjugate actions ($L, G$) as well as the
620: symbols of orbital elements ($a,e$).  However, we should keep in mind that after
621: calculating the average over the fast angles, these actions/elements have the
622: sense of the {\em mean} actions/elements.
623: %
624: %-------------------------------------------------------------------------------
625: \subsection{The effect of quadrupole moment of the star}
626: %-------------------------------------------------------------------------------
627: %
628: To average out the quadrupole moment of the star, we have to calculate integrals
629: from the astrocentric distance of the planet taken in the power of ``-3'' for the spin
630: distortion, and the power of ``-6'' for the tidal distortion. To calculate these averages,
631: we express the astrocentric distance $r$ of the inner planet with respect to the
632: true anomaly and  we replace the integration variables $d\mathcal{M} =
633: \mathcal{J} df$. For $l>1$, we  find
634: \begin{equation}
635: \left<\frac{1}{r^l}\right> = \frac{(1 - e^2)^{1/2}}{a^l 
636: \left(1 - e^2\right)^{l-1}}
637: \sum_{s=0}^{l-2} {{{l-2}\choose{s}} e^s \left<\cos^s{f}\right>},
638: \end{equation}
639: where, for $s$ even, the average of $\left<\cos^s{f}\right>$  over the mean
640: anomaly reads as follows:
641: \begin{equation}
642: \left<\cos^s{f}\right> = \frac{2^s \pi }{\Gamma \left(\frac{1}{2}-\frac{s}{2}\right)^2
643: \Gamma (s+1)},
644: \end{equation}
645: while for $s$ odd, the average over the mean anomaly
646: $\left<\cos^s{f}\right>=0$.  The average of the leading term in
647: $\left<\HflatS\right>$ has the form of:
648: \begin{equation}
649: \left<\HflatS\right> \approx -\frac{\beta  k_L
650:     \left(1 + m_1/m_0\right) R_0^5 \Omega^2}{6 a^3
651:     \left(1-e^2\right)^{3/2}}.
652: \label{avHflat}
653: \end{equation}
654: Having this secular Hamiltonian, we can calculate the apsidal 
655: frequency forced by the spin-induced QM of the star:
656: \begin{equation}
657: \dot{\varpi}_{\idm{spin}} = 
658: \frac{\partial\,\left<\HflatS\right>}{\partial\,G} \approx 
659: \frac{\beta^7 k_L (1 + m_1/m_0) \mu ^3 R_0^5 \Omega^2}{2 G^4 L^3}.
660: \label{oqm}
661: \end{equation}
662: 
663: Similarly, we calculate the leading term in $\left<\HflatT\right>$:
664: \begin{equation}
665: \left<\HflatT\right> \approx -\frac{\beta k_L R_0^5 \left(\frac{3}{8} e^4+3 e^2+1\right) k^2
666:     m_1 \left(1+m_1/m_0\right)}{a^6 \left(1-e^2\right)^{9/2}},
667: \label{avHflatT}
668: \end{equation}
669: and the apsidal frequency due to this correction:
670: \begin{equation}
671: \dot{\varpi}_{\idm{TB}} = 
672: \frac{\partial\,\left<\HflatT\right>}{\partial\,G} \approx 
673: \frac{15 \beta^{13} k_L R_0^5 \left(G^4-14 L^2 G^2+21 L^4\right)
674:     m_1 \mu^7}{8 G^{10} L^7 m_0}.
675: \label{oqmT}
676: \end{equation}
677: The tidal bulge induced by a close companion may be 
678: very important in some  configurations.  In fact, the TB effect may be 
679: even dominant
680: over the spin-induced quadrupole moment perturbations. For instance, for $\upsilon$~Andr~b it is
681: $\sim 20$~times as big \citep[see][for details]{Mardling2002,Nagasawa2005}.
682: %
683: %-------------------------------------------------------------------------------
684: \section{Tools to analyze the secular dynamics}
685: %-------------------------------------------------------------------------------
686: %
687: To overview the dynamical influence of the GR and QM  corrections on
688: the point mass NG dynamics, we compare the {\em secular} evolution of the
689: eccentricity of the innermost planet in the sample of detected multiple extrasolar
690: systems. In this experiment, we basically follow \cite{Adams2006a}, however,
691: solving the {\em averaged} equations of motion by means of the numerical
692: integrator. Moreover, we carry out this test to illustrate the different and rich
693: behaviors of  multi-planet systems in the detected sample; a systematic
694: analysis is described in subsequent sections. The results are shown 
695: in Fig.~\ref{fig:fig1}. Each panel in this figure is labeled with the  parent star name.
696: The orbital elements of the selected systems, and the parameters of their parent
697: stars come  from the Jean Schneider  Encyclopedia of Extrasolar
698: Planets\footnote{http://exoplanet.eu}. In this experiment, we fixed 
699: rotational period of the star, $T_{\idm{rot}}=30$~days, 
700: $k_L=0.02$ (hence $J_2 \sim 10^{-7}$--$10^{-6}$).   In some cases, the differences between predictions of the
701: generalized theory and by the classical model are significant, in some other
702: cases both theories give practically the same  outcome.  Looking at the elements
703: of the examined systems, we may conclude that the corrections become very
704: important  for systems with the innermost planet  close to the star, and with
705: other bodies relatively distant.  
706: 
707: In fact, it is already well known \cite[e.g.,][]{Adams2006a} that the additional
708: effects are important when the strength and characteristic time scale of
709: perturbations stemming from the GR and QM become comparable with the secular
710: time-scale of mutual NG interactions.  Then the GR and QM  regarded as {\em
711: perturbing} effects may induce significant  contribution to the secular
712: evolution and the interplay of these effects with the NG interactions may lead
713: to very complex and rich dynamics. As we will see below, this condition is
714: particularly well satisfied  for strictly hierarchical  systems, e.g.,
715: HD~217107, HD~38529 \citep{Fischer2001},  HD~190360~\citep{Vogt2005},
716: HD~11964~\citep{Butler2006},  $\upsilon$ Andromedae~\citep{Butler1999}. The
717: secular time scales can be also comparable when the planetary masses are
718: relatively small because the NG-induced apsidal  frequency decreases
719: proportionally to the products of these masses while, in the first
720: approximation, the GR and spin-induced apsidal motion of the innermost orbit
721: does not depend directly on the planetary mass (see Eqs.~\ref{orel},~\ref{oqm}).
722: 
723: Still, the  results illustrated in Fig.~\ref{fig:fig1}  have limited
724: significance for the study of the global dynamics. Drawing one-dimensional
725: time--orbital  element plots, we can analyze the dynamical evolution only for a
726: few isolated initial conditions. This can be a serious drawback, if we recall
727: that the initial conditions of the discovered systems are still known with large
728: uncertainties. Some  critically important parameters governing the secular
729: evolution, like the masses, nodal lines and inclinations of orbits are poorly
730: constrained by the observations or undetermined at all.  There is also a
731: technical problem: the  direct integrations over the secular time-scale are CPU
732: intensive. Yet looking at  isolated configurations, we obtain only a local view of
733: the dynamics. Instead, following the methodology of Poincar\'e, we should try to
734: understand globally the perturbing effects and the resulting dynamics. We should
735: investigate the whole families of solutions rather than a few isolated
736: phase-space trajectories. In such a case, the application  of secular analytical
737: theories become critically important.
738: \ifpdf
739: \begin{figure*}
740: \centering
741: \includegraphics[width=17.6cm]{fig1.pdf}
742: \caption{
743: The long-term, secular  evolution of eccentricities of the innermost planets of
744: known non-resonant extrasolar systems. Red curves are for the point-mass
745: planets, interacting mutually through Newtonian forces. The blue curves are for 
746: generalized model of motion, including additional effects  (general relativity
747: and the quadrupole moment of the star). For all  tested planetary systems, we
748: fix $T_{\idm{rot}} = 30$~days, $k_L=0.02$. The names of  parent stars are
749: labeling each panel together with  a number of planets written in brackets.
750: }
751: \label{fig:fig1}
752: \end{figure*}
753: \else
754: \begin{figure*}
755: \centering
756: \includegraphics[width=17.6cm]{fig1.eps}
757: \caption{
758: The long-term, secular  evolution of eccentricities of the innermost planets of
759: known non-resonant extrasolar systems. Red curves are for the point-mass
760: planets, interacting mutually through Newtonian forces. The blue curves are for 
761: generalized model of motion, including additional effects  (general relativity
762: and the quadrupole moment of the star). For all  tested planetary systems, we
763: fix $T_{\idm{rot}} = 30$~days, $k_L=0.02$. The names of  parent stars are
764: labeling each panel together with  a number of planets written in brackets.
765: }
766: \label{fig:fig1}
767: \end{figure*}
768: \fi
769: %
770: %-------------------------------------------------------------------------------
771: \subsection{Representative plane of initial conditions and equilibria}
772: %-------------------------------------------------------------------------------
773: %
774: The simplest class of solutions which can be studied most effectively, are the
775: equilibria (or stationary solutions).  In the multi-parameter dynamical systems,
776: their positions (coordinates), stability, and bifurcations provide  information
777: on the general structure of the phase space. Hence, we focus on the
778: stationary solutions emerging in the secular model of a two-planet system with
779: the GR and QM corrections. 
780: 
781: After the averaging of $\Hfull$ over the mean anomalies (i.e., over
782: the orbital periods), the secular Hamiltonian ($\Hsec$) does not depend on $l_i
783: \equiv \mathcal{M}_i$ anymore. Therefore, as we mentioned already,
784: the conjugate actions $L_i$ become
785: constants of motion (hence, the mean semi-major axes are also constant).
786: Moreover, in the coplanar problem the longitudes of nodes are undefined and
787: irrelevant for the dynamics. The secular energy of two-planet system does not
788: depend on individual longitudes of pericenters $\varpi_i$, but only on the their
789: difference  $\Delta\varpi = \varpi_1-\varpi_2$ \citep{Brouwer1961}. These basic
790: facts can be expressed  with the help of appropriate canonical transformation
791: \citep{Michtchenko2004}:
792: \begin{eqnarray}
793: \label{eqtot1}
794: \Delta\varpi \equiv \varpi_1 - \varpi_2, \quad &&G_1,\\
795: \label{eqtot2}
796: \varpi_2, \quad &&C = G_1 + G_2.
797: \end{eqnarray}
798: Because $\varpi_2$ is the cyclic angle, the conjugate momentum equal to the
799: total angular momentum $C$ is conserved. For a fixed 
800: angular momentum as a constant parameter, the  phase-space  of reduced system
801: become two-dimensional,  and  the system is integrable.  We can now choose
802: $\Delta\varpi$ as the canonical angle and $G_1$ as the
803: conjugate momentum. Alternatively, the role of the momentum may be attributed to
804: $e_1$. Obviously,  $G_2$ (or $e_2$) becomes a dependent
805: variable (through the $C$ integral).  
806: 
807: To study the dynamics of the reduced secular system in a global manner, we
808: follow a concept of the so called  {\em representative plane} of initial
809: conditions introduced by \cite{Michtchenko2004}. The representative plane
810: (\RP{}) comprises of points in the phase space which  lie on a specific plane
811: crossing all phase trajectories of the system. In the cited paper, it has been
812: shown that a good choice of the \RP{}-plane flows from the condition of
813: vanishing derivatives of the secular Hamiltonian over $\Delta{\varpi}$. It is
814: equivalent to the symmetry of secular interactions with respect to fixed apsidal
815: line of a selected orbit and it means that time derivatives of the conjugate
816: momenta (which can be expressed by eccentricities) must vanish.  Indeed, in
817: accord with the law of conservation of the  total angular momentum,  the
818: eccentricities must reach at the same time maximal and minimal values along the phase
819: trajectories, hence $\dot{e_i}=0$, $i=1,2$.  Regarding the classic problem with
820: Newtonian point-to-point interactions, this condition is satisfied when
821: $\Delta{\varpi}=0$ or $\Delta{\varpi}=\pi$ \citep{Michtchenko2004} [see also
822: \citep{Migaszewski2008a}]. If we consider the generalized problem of two-planet
823: system then the concept of the representative plane is still valid. The {\em
824: small} perturbations which we introduce do not change the dimension of the phase
825: space and the motion remains on perturbed Keplerian orbits, only the form of the
826: secular Hamiltonian is modified. Moreover, we may expect that qualitative
827: properties of the secular system may be changed because we introduce 
828: a few new
829: parameters describing physical setup of the  studied system and interactions
830: governing its dynamical evolution. To encompass both the specific values of
831: angle $\Delta{\varpi}=0,\pi$, the representative plane can be defined by the
832: following set of points:
833: \[
834: {\cal S} = \{ e_1 \cos{\Delta{\varpi}} \times e_2;  e_1, e_2 \in [0,1) \}, 
835: \]
836: where $\Delta\varpi=\pi$,  $e_1\cos{\Delta{\varpi}}<0$ for the left-hand
837: half-plane (called the \RPm-plane, from hereafter), and $\Delta\varpi=0$,
838: $e_1\cos{\Delta{\varpi}}>0$ are for the right-hand half-plane (the \RPp-plane).
839: Hence, the sign of $e_1 \cos\Delta\varpi$ on the $x$-axis tells us on the value of angle
840: $\Delta\varpi$.
841: 
842: \subsection{A general view of the representative plane}
843: To illustrate the influence of additional perturbations on the secular dynamics
844: of the classic model, we compute a set of the \RP{}-planes for wide ranges of
845: orbital and physical parameters of the studied two-planet  configurations.
846: Before we discuss these results, at first  we explain how the \RP-{plane}
847: illustrates  the dynamical structure of the phase space.  Conveniently,
848: Figure~\ref{fig:fig2}, which is derived for {\em specific} orbital parameters
849: (given in the caption), reveals  {\em all} relevant features in a condensed form
850: and its description can be regarded as a guide useful for the further analysis.
851: 
852: We start from the left-hand panel of Fig.~\ref{fig:fig2}. Smooth half-ellipse
853: like curves marked with thin lines 
854: are for the levels of the secular energy of the
855: generalized problem. The straight lines  are for the collision line of orbits
856: defined through  $a_1 (1\pm e_1) = e_2 (1-e_2)$.
857: %
858: \ifpdf
859: \begin{figure*}
860: \centerline{
861: \hbox{\includegraphics[width=8.8cm]{fig2a.pdf}
862:       \hspace*{0cm}   
863:       \includegraphics[width=8.8cm]{fig2b.pdf}
864: }
865: }
866: \caption{
867: The representative plane $(e_1 \cos{\Delta{\varpi}}, e_2)$ for the two-planet
868: system. The elements are $m_1=1.359~\mbox{m}_{\idm{J}}$,
869: $m_2=0.453~\mbox{m}_{\idm{J}}$, $a_1=0.1$~au, $a_2=1.0$~au. Stellar mass is
870: 1~$M_{\sun}$, rotational period is $T_{\idm{rot}}=30$~days, stellar radius is
871: 1~$R_{\sun}$, $k_L=0.02$. The thick curves are for the stationary solutions: red
872: curves are for stable equilibria in the classic model, blue and violet curves
873: are for stable and unstable equilibria of the generalized model, respectively.  
874: {\em The left panel}:  Color contours are for the ratio of the  apsidal
875: frequency induced by the general relativity and quadrupole moment  to the 
876: apsidal frequency caused by mutual interactions between planets, $\kappa$.
877: \corr{The scale of color code is limited by $\kappa=1$, meaning that the
878: corrections become more important than the point mass Newtonian gravity. In
879: general, the yellow colour may refer to $\kappa>1$.} Energy levels of the
880: generalized model are marked with thin lines.  {\em The right panel}: A few
881: particular energy levels of the generalized model are labeled with
882: \textbf{a}---\textbf{f}, accordingly.  See also phase diagrams for these energy
883: levels  which are shown in Fig.~\ref{fig:fig3} and Fig.~\ref{fig:fig4}. Curves
884: of equilibria are labeled: mode~I are for aligned apsides ($\Delta\varpi=0$),
885: mode~II are for anti-aligned apsides ($\Delta\varpi=\pi$),  \UE{} means unstable
886: equilibria.  The thin half-ellipse like curves are for energy levels computed
887: for the classic model (red) and generalized model (green). More details can be
888: found in the text. 
889: }
890: \label{fig:fig2}
891: \end{figure*}
892: \else 
893: \begin{figure*}
894: \centerline{
895: \hbox{\includegraphics[width=8.8cm]{fig2a.eps}
896:       \hspace*{0cm}   
897:       \includegraphics[width=8.8cm]{fig2b.eps}
898: }
899: }
900: \caption{
901: The representative plane $(e_1 \cos{\Delta{\varpi}}, e_2)$ for the two-planet
902: system. The elements are $m_1=1.359~\mbox{m}_{\idm{J}}$,
903: $m_2=0.453~\mbox{m}_{\idm{J}}$, $a_1=0.1$~au, $a_2=1.0$~au. Stellar mass is
904: 1~$M_{\sun}$, rotational period is $T_{\idm{rot}}=30$~days, stellar radius is
905: 1~$R_{\sun}$, $k_L=0.02$. The thick curves are for the stationary solutions: red
906: curves are for stable equilibria in the classic model, blue and violet curves
907: are for stable and unstable equilibria of the generalized model, respectively.  
908: {\em The left panel}:  Color contours are for the ratio of the  apsidal
909: frequency induced by the general relativity and quadrupole moment  to the 
910: apsidal frequency caused by mutual interactions between planets, $\kappa$.
911: \corr{The scale of color code is limited by $\kappa=1$, meaning that the
912: corrections become more important than the point mass Newtonian gravity. In
913: general, the yellow colour may refer to $\kappa>1$.} Energy levels of the
914: generalized model are marked with thin lines.  {\em The right panel}: A few
915: particular energy levels of the generalized model are labeled with
916: \textbf{a}---\textbf{f}, accordingly.  See also phase diagrams for these energy
917: levels  which are shown in Fig.~\ref{fig:fig3} and Fig.~\ref{fig:fig4}. Curves
918: of equilibria are labeled: mode~I are for aligned apsides ($\Delta\varpi=0$),
919: mode~II are for anti-aligned apsides ($\Delta\varpi=\pi$),  \UE{} means unstable
920: equilibria.  The thin half-ellipse like curves are for energy levels computed
921: for the classic model (red) and generalized model (green). More details can be
922: found in the text. 
923: }
924: \label{fig:fig2}
925: \end{figure*} 
926: \fi
927: %
928: The relative magnitude of the corrections to the secular Hamiltonian
929: is represented by contour levels of the following coefficient:
930: \begin{equation}
931: \label{eq:kappa}
932: \kappa(e_1 \cos \Delta\varpi,e_2)
933: = \left|\frac{\partial\left<\Hrel\right>/\partial{G_1} +
934: \partial\left<\Hflat\right>/\partial{G_1}}
935: {\partial\left<\Hmut\right>/\partial{G_1}}\right|,
936: \end{equation}
937: i.e., the ratio of the apsidal frequency of the inner pericenter induced by
938: $\left<\Hrel + \Hflat\right>$ relative to the  ``natural''  apsidal frequency in
939: the point mass Newtonian model. Regions of the \RP{}-plane  where $\kappa>0.1$, are
940: color-coded, according to the levels of constant $\kappa$. Yellow color (light
941: gray in the printed paper) encodes $\kappa \ge 1$, meaning that
942:  the apsidal frequency of
943: the innermost pericenter forced by  {\em perturbing} GR+QM corrections is
944: larger than the relative pericenter frequency $\dot{\Delta\varpi}$ 
945: caused by the secular NG interactions between  planets. 
946: 
947: The thick curves defined through:
948: \begin{equation}
949: \frac{\partial{\Hsec}}{\partial{G_1}} = 0,
950: \end{equation}
951: can be attributed to the stationary solutions of the reduced, two-dimensional system
952: with $(G_1,\Delta\varpi)$-variables, because  coordinates of all points of 
953: these curves must also satisfy
954: \begin{equation}
955: \frac{\partial{\Hsec}}{\partial{\Delta{\varpi}}} = 0,
956: \end{equation}
957: according to the definition of the \RP{}-plane.  Such solutions are periodic
958: orbits of the full secular system (Eqs.~\ref{eqtot1}--\ref{eqtot2}). We  examine
959: the Lyapunov stability of these equilibria with the help of the Lyapunov
960: theorem, adopting  $\Hsec$ as the Lyapunov function in the cases when they 
961: correspond to its maximum (or a minimum) in the reduced two-dimensional phase
962: space.  \corr{Because the investigated dynamical system has one degree of
963: freedom, the stable or unstable equilibria can be easily identified as extrema
964: or saddles of the secular Hamiltonian in the representative plane,
965: respectively}. 
966: 
967: The positions of equilibria [or stationary modes \citep{Michtchenko2004}] help
968: us to distinguish between different types (families) of orbits characterized by
969: librations of $\Delta\varpi$ around a particular value (libration center).  The
970: red, thick curves drawn in both panels of Fig.~\ref{fig:fig2} are for stationary
971: modes in the classic model \citep{Michtchenko2004}, while the equilibria in the
972: generalized model are drawn with thick, blue curves. In the \RPp{} half-plane,
973: these stationary solutions are classified by \cite{Michtchenko2004} as mode~I
974: equilibria, and they are characterized  by librations of $\Delta\varpi$
975: around~$0$ in the neighboring trajectories.  In the \RPm{} half-plane, we can
976: find also  Lyapunov stable mode~II solutions related to librations of
977: $\Delta\varpi$ around~$\pi$ in close trajectories.  Some parts of the equilibria
978: curves are marked with   violet color (for the generalized model). These points
979: denote unstable equilibria (\UE{}) accompanied by {\em the true secular
980: resonance} (the \TSR{} from hereafter). Such unstable equilibria in the
981: \RPp{}-plane  are discovered  by \cite{Michtchenko2004} in the secular  classic
982: coplanar model of 2-planets. Obviously, mode~I and mode~II solutions known as
983: generic features of the classic model, exist also in the generalized problem.
984: However, their  positions in the phase space may be heavily affected by
985: apparently subtle GR and QM perturbations.  
986: 
987: Note that along the red thick curves representing equilibria in the classic
988: model, $\kappa$ is undefined. 
989: 
990: Now, let us examine the right-hand panel of Fig.~\ref{fig:fig2}. In this plot,
991: besides levels of the secular energy of the classic model (red, thin curves), we
992: also plot  such levels for the generalized problem (green, thin curves). We can
993: observe a significant discrepancy between the shapes of contour levels of both
994: Hamiltonians. 
995: 
996: In this plot we mark also a few specific levels of $\Hsec$: 
997: $\mathcal{E}_{\idm{\bf a}} = -2.247$,
998: $\mathcal{E}_{\idm{\bf b}} = -2.24955$,
999: $\mathcal{E}_{\idm{\bf c}} = -2.25$,
1000: $\mathcal{E}_{\idm{\bf d}} = -2.2525$,
1001: $\mathcal{E}_{\idm{\bf e}} = -2.256$,
1002: $\mathcal{E}_{\idm{\bf f}} = -2.26$,
1003: respectively; here, we skipped all constant terms (the Keplerian part, indirect
1004: term, constant relativistic term) in the full secular Hamiltonian, and  the
1005: energy values are given in terms of $10^{-5}~M_{\sun}\mbox{au}^2\mbox{yr}^{-2}$ 
1006: when 1~$M_{\sun}$, 1~au, and 1~sideral year are taken as units of mass,
1007: distance, and time, respectively (then the Gauss constant $k=2\pi$).  Because
1008: the motion of the secular system  must be confined to fixed energy level, points
1009: at which a particular energy level crosses the curve representing stationary
1010: solutions, tell us on distinct equilibria and their bifurcations. For instance,
1011: following energy level labeled with {\bf a} in the right-hand panel of Fig.~2,
1012: we see that it  crosses the equilibria curves at two points. The first one,
1013: found in the \RPp{} half-plane, corresponds to mode~I solution.  The second
1014: cross-point is found in the \RPm{} half-plane, and it marks the mode~II
1015: equilibrium.
1016: 
1017: To estimate the precision of the analytical theory, the energy levels  in
1018: Fig.~\ref{fig:fig2} are calculated with exact semi-analytical averaging
1019: \citep{Michtchenko2004} that makes use of precise adaptive Gauss-Legendre
1020: quadratures \citep{Migaszewski2008b}.  For a comparison, the stationary modes
1021: are computed with both methods: we recall that thick curves represent equilibria
1022: calculated with the semi-numerical averaging while the thin curves (over-plotted
1023: on them) are for  stationary solutions calculated with the help of the
1024: analytical theory outlined in Sect.~2 (the right-hand panel of
1025: Fig.~\ref{fig:fig2}). The results of these methods are in excellent accord. The
1026: most significant  discrepancies between the results appear in the \RPp{}
1027: half-plane (for $\Delta\varpi=0$), in the regime of large eccentricity $e_2$. In
1028: such a case, the secular series representing $\Hmut$ diverge over the {\em
1029: anti-collision} line (marked with red, straight line). This problem is discussed
1030: in detail in \citep{Migaszewski2008a}.  Obviously, over this anti-collision
1031: line, $r_2>r_1$ at some parts of orbits, breaking the underlying assumption that
1032: we require to expand the inverse of the mutual distance in convergent series.
1033: Moreover, we demonstrate that the analytical theory reproduces the dynamics of 
1034: hierarchical configurations up to very large $e_2$. Another, direct test of the
1035: precision of the analytic theory is given in Sect.~3.4.
1036: %
1037: %-------------------------------------------------------------------------------
1038: \subsection{Phase diagrams and the non-linear secular resonance}
1039: %-------------------------------------------------------------------------------
1040: %
1041: To investigate more closely the secular dynamics related to the new mode~II and
1042: \UE{} solutions, and to understand the structure of the phase space in more
1043: details, we compute a number of {\em phase plots} (or {\em phase
1044: diagrams}). The secular energy is kept constant and we draw the phase
1045: trajectories in the ($e_1 \cos{\Delta{\varpi}}, e_1 \sin{\Delta{\varpi}}$)--, as
1046: well as  ($e_2 \cos{\Delta{\varpi}}, e_2 \sin{\Delta{\varpi}}$)--planes,
1047: choosing the initial conditions along the fixed energy level.  The phase plots
1048: are computed with the help of the analytic theory.  For a reference, we recall
1049: the right-hand panel of Fig.~\ref{fig:fig2} which illustrates the  \RP{}-plane
1050: of $(e_1 \cos{\Delta{\varpi}}, e_2)$.   We recall that the \UE{} of the
1051: generalized model are marked with violet curves. 
1052: 
1053: 
1054: The phase diagrams are illustrated in Fig.~\ref{fig:fig3}. The top row is for
1055: the energy level of $\mathcal{E}_{\idm{a}} = -2.247\times10^{-5}$  which is
1056: labeled with {\bf a} in the right-hand panel of Fig.~\ref{fig:fig2}.  This
1057: energy level crosses the equilibria curves in two points which  can be
1058: recognized in the phase diagrams as libration centers of $\Delta\varpi=0$
1059: (labeled with mode~I), and of $\Delta\varpi=\pi$  (labeled with mode~II),
1060: respectively. Overall, the phase diagrams look like qualitatively the same as in
1061: the classic model. Both libration modes are separated from the circulation zone
1062: of $\Delta\varpi$ by {\em false} separatrices
1063: \citep{Pauwels1983,Michtchenko2004}, i.e., the transition between each mode and
1064: the circulation  of $\Delta\varpi$ {\em does not} involve solutions with
1065: infinite period. \corr{ This is illustrated in smaller, bottom plots
1066: accompanying each phase diagram, which show the secular, fundamental frequency
1067: $g$ of the mean system for initial conditions lying on the $x$ axis of the
1068: respective phase diagram.   This frequency has been determined through 
1069: solutions of the secular equations of motion. Clearly, when the true separatrix
1070: crosses the $e_{1,2}\cos\Delta\varpi$-axis, $g$ decreases to 0. Crossings of
1071: false separatrices do not lead to any discontinuity of smooth plots of $g$. }
1072: %
1073: \ifpdf
1074: \begin{figure*}
1075: \centerline{
1076: \vbox{
1077: \hbox{\includegraphics[width=68mm]{fig3a.pdf}
1078: \hspace*{6mm}
1079:       \includegraphics[width=68mm]{fig3b.pdf}}    
1080: \hbox{\includegraphics[width=68mm]{fig3c.pdf}
1081: \hspace*{6mm}
1082:       \includegraphics[width=68mm]{fig3d.pdf}}
1083: }
1084: }      
1085: \caption{
1086: Phase diagrams computed for two-planet secular system described with the same
1087: parameters as used for the construction of Fig. \ref{fig:fig2}. \corr{ Each
1088: phase diagram is accompanied by a smaller  plot of the 
1089: fundamental frequency $g$ of
1090: the secular solutions  calculated for initial conditions lying on  the $y \equiv
1091: e_{1,2}\cos\Delta\varpi$-axis. } Panels in the top row are for the secular
1092: energy $E_{\bf a} = -2.247\times10^{-5}$, panels in the bottom row are for
1093: $E_{\bf c} = -2.25\times10^{-5}$. These levels are labeled in Fig.
1094: \ref{fig:fig2} with ({\bf a}) and ({\bf c}), respectively.  The left-hand panels
1095: are for the ($e_1 \cos{\Delta{\varpi}}, e_1 \sin{\Delta{\varpi}}$)-plane, the
1096: right-hand panels are  for ($e_2 \cos{\Delta{\varpi}}, e_2
1097: \sin{\Delta{\varpi}}$)-plane. Shaded regions mark different  libration zones
1098: around mode~I ($\Delta\varpi=0$) and mode~II ($\Delta\varpi=\pi$), respectively.
1099: \UE{} marks  unstable equilibria  accompanied by libration zones around
1100: $\Delta\varpi=\pi$ and centered at the \TSR{} solutions.  \corr{The true
1101: separatrices are indicated by $g \rightarrow 0$}. See the text for more details.
1102: }
1103: \label{fig:fig3}
1104: \end{figure*}
1105: \else
1106: \begin{figure*}
1107: \centerline{
1108: \vbox{
1109: \hbox{\includegraphics[width=68mm]{fig3a.eps}
1110: \hspace*{6mm}
1111:       \includegraphics[width=68mm]{fig3b.eps}}    
1112: \hbox{\includegraphics[width=68mm]{fig3c.eps}
1113: \hspace*{6mm}
1114:       \includegraphics[width=68mm]{fig3d.eps}}
1115: }
1116: }      
1117: \caption{
1118: Phase diagrams computed for two-planet secular system described with the same
1119: parameters as used for the construction of Fig. \ref{fig:fig2}. \corr{ Each
1120: phase diagram is accompanied by a smaller  plot of the 
1121: fundamental frequency $g$ of
1122: the secular solutions  calculated for initial conditions lying on  the $y \equiv
1123: e_{1,2}\cos\Delta\varpi$-axis. } Panels in the top row are for the secular
1124: energy $E_{\bf a} = -2.247\times10^{-5}$, panels in the bottom row are for
1125: $E_{\bf c} = -2.25\times10^{-5}$. These levels are labeled in Fig.
1126: \ref{fig:fig2} with ({\bf a}) and ({\bf c}), respectively.  The left-hand panels
1127: are for the ($e_1 \cos{\Delta{\varpi}}, e_1 \sin{\Delta{\varpi}}$)-plane, the
1128: right-hand panels are  for ($e_2 \cos{\Delta{\varpi}}, e_2
1129: \sin{\Delta{\varpi}}$)-plane. Shaded regions mark different  libration zones
1130: around mode~I ($\Delta\varpi=0$) and mode~II ($\Delta\varpi=\pi$), respectively.
1131: \UE{} marks  unstable equilibria  accompanied by libration zones around
1132: $\Delta\varpi=\pi$ and centered at the \TSR{} solutions.  \corr{The true
1133: separatrices are indicated by $g \rightarrow 0$}. See the text for more details.
1134: }
1135: \label{fig:fig3}
1136: \end{figure*}
1137: \fi
1138: %
1139: The phase diagrams are much more complicated for energy levels labeled with {\bf
1140: b}--{\bf f}, which cross the equilibria curves at more than two points. We
1141: analyze in detail the energy level ({\bf c})  which intersects the curve of
1142: stationary solutions in {\em four} points. These points can be recognized in the
1143: phase diagrams shown in the bottom panels of Fig.~\ref{fig:fig3}. We can
1144: identify them easily in the $x$-axis of these diagrams  (because,  such points
1145: of the \RP{}-plane have $y \equiv  e_{1,2}\sin \Delta\varpi=0$). Starting at the
1146: \RPp{}-plane and following the energy level counter-clockwise, we have a stable
1147: mode~I equilibrium surrounded by large zone of librations of $\Delta\varpi$
1148: around~$0$ which corresponds to single crossing point of the energy level and
1149: the equilibria curve in the \RPp{}-plane. In the \RPm{}-plane, we have {\em
1150: three} such points, two of them are Lyapunov stable, and one point in the middle
1151: is an unstable equilibrium. This part of the phase space, as seen in
1152: Fig.~\ref{fig:fig3}, encompass figure-eight shaded area, involving two islands
1153: of librations (\TSR{}s, or elliptic points) around $\Delta\varpi=\pi$,  and the
1154: hyperbolic \UE{} point lying in the middle between them. Both libration centers
1155: are characterized by $\Delta\varpi=\pi$. Because they are related to stable
1156: equilibria separated by hyperbolic structure of the \UE{}, two parts of  the
1157: phase curve surrounding the islands of the \TSR{}s and that meet in the \UE{},
1158: must form a real separatrix. The whole structure may be  still surrounded by a
1159: zone of librations of $\Delta\varpi$ around $\pi$. It is shaded in light-gray.
1160: Let us note, that this mode~II libration area is confined to the \RPm{}-plane
1161: (hence, in this particular case, angle $\Delta\varpi$  does not pass through
1162: $0$).
1163: 
1164: A sequence of ($e_2 \cos{\Delta{\varpi}}, e_2 \sin{\Delta{\varpi}}$)-diagrams
1165: for the secular energy levels {\bf a}--{\bf f} are shown in Fig.~\ref{fig:fig4}.
1166: Looking at these levels plotted in the \RP{}-plane, we can now follow a
1167: development of dynamical structures related to the different modes of motion. In
1168: particular, phase diagrams Fig.~\ref{fig:fig4}b and Fig.~\ref{fig:fig4}d reveal
1169: bifurcations of mode~II which emerge the \UE{} and  \TSR{} solutions. Clearly,
1170: the bifurcations may  be identified with points at which the given energy level
1171: is tangent to the equilibria curves in the \RP{}-plane. 
1172: 
1173: The phase diagrams assure us that the secular dynamics of the generalized model
1174: can be much more complex and rich due to the GR and QM corrections to the NG
1175: Hamiltonian than the secular dynamics in the classic model.
1176: %
1177: %-------------------------------------------------------------------------------
1178: \subsection{Numerical test of the analytic secular theory}
1179: %-------------------------------------------------------------------------------
1180: %
1181: Finally, we illustrate limitations of  the secular theory and we compare its
1182: results with the outcome of the direct numerical integrations. This comparison
1183: is also directly related to the dynamical stability of the planetary system.
1184: 
1185: First, we constructed the numerical model of the generalized system
1186: independently on the analytical model. We wrote the equations of motion with
1187: respect to the Jacobi reference frame, using formulation of \cite{Mardling2002} 
1188: who call it {\em the direct code}. In this code, the GR acceleration is modeled 
1189: with the PPN formulae given in \citep{Kidder1995}. In that way we have a
1190: possibility to check the analytic theory in completely independent way, which
1191: also prevents copying logical errors which could be done during the averaging.
1192: After some experiments, we also found that the choice of the reference frame
1193: (e.g., related to  Jacobi, Poincar\'e, or classic-astrocentric coordinates) is
1194: in fact irrelevant for the results of this test. We also do not
1195: account for the difference between the {\em osculating} and the {\em mean}
1196: elements. 
1197: 
1198: Using the direct code, we integrated numerically a few phase diagrams 
1199: for  nominally {\em
1200: the same initial conditions and parameters} used to draw plots in
1201: Fig.~\ref{fig:fig4}a--f with the help of the analytic, secular model. We computed osculating elements related
1202: to the Jacobi reference  frame over a few secular cycles. The numerically
1203: derived phase curves are drawn with black, filled circles in  respective panels
1204: of Fig.~\ref{fig:fig4}. The solutions obtained with the analytic theory are
1205: over-plotted on these numerical solutions with thiner, green curves. Subsequent
1206: five panels  of Fig.~\ref{fig:fig4}a--e reveal that the agreement of both sets
1207: of solutions is excellent. The analytic theory reproduces qualitative features
1208: seen in the phase plots, and their structure with great accuracy. We find that
1209: both solutions coincide even in the regime of large eccentricities.
1210: Remarkably, the direct code integrations last over CPU time which is by a few
1211: orders of magnitude longer than the calculations carried out with the help of
1212: the analytical theory.
1213: 
1214: However, in the last panel of Fig.~\ref{fig:fig4}f we can observe significant
1215: deviations of the analytic solutions from the numerical theory, particularly in
1216: the outer parts of the phase diagram.  In fact, in this case $e_2$ is so large that
1217: the assumptions of the secular theory are broken. After examining the
1218: \RP{}-plane (Fig.~\ref{fig:fig2}), we can see that the energy level
1219: corresponding to the last panel passes close to the collision line. To 
1220: illustrate the real border of the dynamical stability, we examined the dynamical
1221: character of solutions in the \RP{}-plane with the help of the Spectral Number
1222: technique \citep{Michtchenko2001}. This simple FFT-based algorithm makes it
1223: possible to distinguish between chaotic and regular solutions. The dynamical
1224: maps shown in Fig.~\ref{fig:fig5} are constructed by counting the number of
1225: frequencies in the FFT-spectrum of the time series, $\{ \sigma(t) = a(t) \exp
1226: \mbox{i}\lambda_i(t) \}$, where $a_i(t)$ and $\lambda_i(t)$ are temporal {\em
1227: canonical} semi-major axes and mean longitude of each planet. The number of
1228: peaks (the Spectral Number, SN from hereafter)  in the spectrum over some noise
1229: level tells us on the character of orbit. Orbits with large SN (grater than
1230: 1000) are very chaotic, while the SN $\sim 1$ means small number of frequencies
1231: and a regular, quasi-periodic phase trajectory. Each point in the dynamical maps
1232: represents a phase trajectory  that was integrated over $\sim 10^4~P_{\idm{2}}$.
1233: Although such time span is relevant for the short-term dynamics only, 
1234: calculations took a very long CPU time (a few days on 24 AMD-CPU cores). The
1235: results are shown in two panels of Fig.~\ref{fig:fig5}. The left panel is for
1236: the classic model (only NG interactions are included), while the right panel is
1237: constructed for the generalized model.
1238: 
1239: Let us analyze the left-hand panel of Fig.~\ref{fig:fig5}
1240:  for the classic model. Solutions, which appear
1241: strongly chaotic  are marked with colors (darker point means larger SN and more
1242: chaotic system).  Clearly, the border of stable motions is irregular and is
1243: shifted towards small $e_2$ by $0.1$--$0.2$ with respect to the formal,
1244: geometrical collision line bordering the triangular region (it is drawn in both
1245: panels of Fig.~\ref{fig:fig5}). The thick red curves mark the equilibria of mode
1246: I and mode~II, respectively. Shaded regions are for the initial conditions in
1247: the \RP{}-plane corresponding to orbital configurations with librating
1248: $\Delta\varpi$.  In the right-hand panel of Fig.~\ref{fig:fig5}, we show the
1249: equilibria curves of  the generalized model and  libration zones of
1250: $\Delta\varpi$ associated with these equilibria. We mark again the SN signature
1251: of  the short-term dynamics. We note that the border of stability is
1252: quite different from that ones of the classic model (compare with the
1253: left-hand panel of Fig.~\ref{fig:fig5}).  This is a very clear example showing
1254: that apparently subtle GR+QM effects may affect the {\em short-term} stability
1255: of the system in a significant way. 
1256: 
1257: The dynamical map for the generalized model 
1258: (Fig.~\ref{fig:fig5}, the right hand panel) helps us to identify the source of
1259: unstable behavior seen in Fig.~\ref{fig:fig4}f,  revealing that some initial
1260: conditions lead to erratic and irregular behavior. In the dynamical map, we mark
1261: two levels of $\Hsec$ corresponding to phase diagrams drawn in
1262: Fig.~\ref{fig:fig4}e,f, respectively. The energy level ({\bf e}) lies entirely
1263: in the regular region, in spite that $e_2$ may reach values as large as 0.8. The
1264: neighboring level ({\bf f}) may touch the unstable zone, and that is why the
1265: orbital evolution at this energy level  may become very unstable.  Because the
1266: motion must be confined to the fixed energy level, due to the secular evolution,
1267: some initial conditions may be transported to the chaotic zone  (by excitation
1268: of the eccentricity) during  roughly a half of the secular period. Then the
1269: short-term, strong chaos can destabilize the system immediately.  Following the
1270: fixed energy curve, we can also identify three islands of stable motions. The
1271: first one lies in the right half-plane of the phase diagram and is associated
1272: with librations of $\Delta\varpi$ around $0$, see the neighborhood of the
1273: corresponding  cross point in the \RPp-plane, Fig.~\ref{fig:fig2}.  In the
1274: \RPm-plane, we can find corresponding unstable equilibrium and two stable
1275: solutions with associated libration islands shown in Fig.~\ref{fig:fig4}f (see
1276: the left-hand half-plane of the phase diagram) around $(0,-0.1)$ and
1277: $(0,-0.78)$, respectively. 
1278: %
1279: \ifpdf
1280: \begin{figure*}
1281: \centerline{
1282: \vbox{
1283: \hbox{\includegraphics[width=58mm]{fig4a.png}
1284:       \includegraphics[width=58mm]{fig4b.png}
1285:       \includegraphics[width=58mm]{fig4c.png}
1286:       }
1287: \hbox{\includegraphics[width=58mm]{fig4d.png}
1288:       \includegraphics[width=58mm]{fig4e.png}
1289:       \includegraphics[width=58mm]{fig4f.png}
1290:       }
1291: }
1292: }      
1293: \caption{
1294: Phase diagrams in the $(e_2\cos{\Delta{\varpi}}, e_2
1295: \sin{\Delta{\varpi}})$-plane drawn  to illustrate a comparison of the secular
1296: evolution of the mean system (green, thin curves) with the numerical solutions
1297: of full,  unaveraged system (larger, filled black circles). The orbital
1298: parameters are the same as in Fig. \ref{fig:fig2}. Subsequent panels labeled
1299: with \textbf{a}-\textbf{f}  correspond to  relevant energy levels marked and
1300: labeled in Fig.~\ref{fig:fig2}, accordingly.
1301: }
1302: \label{fig:fig4}
1303: \end{figure*}
1304: \else
1305: \begin{figure*}
1306: \centerline{
1307: \vbox{
1308: \hbox{\includegraphics[width=58mm]{fig4a.eps}
1309:       \includegraphics[width=58mm]{fig4b.eps}
1310:       \includegraphics[width=58mm]{fig4c.eps}
1311:       }
1312: \hbox{\includegraphics[width=58mm]{fig4d.eps}
1313:       \includegraphics[width=58mm]{fig4e.eps}
1314:       \includegraphics[width=58mm]{fig4f.eps}
1315:       }
1316: }
1317: }      
1318: \caption{
1319: Phase diagrams in the $(e_2\cos{\Delta{\varpi}}, e_2
1320: \sin{\Delta{\varpi}})$-plane drawn  to illustrate a comparison of the secular
1321: evolution of the mean system (green, thin curves) with the numerical solutions
1322: of full,  unaveraged system (larger, filled black circles). The orbital
1323: parameters are the same as in Fig. \ref{fig:fig2}. Subsequent panels labeled
1324: with \textbf{a}-\textbf{f}  correspond to  relevant energy levels marked and
1325: labeled in Fig.~\ref{fig:fig2}, accordingly.
1326: }
1327: \label{fig:fig4}
1328: \end{figure*}
1329: \fi
1330: %
1331: \ifpdf
1332: \begin{figure*}
1333: \centerline{
1334: \hbox{\includegraphics[width=8.7cm]{fig5a.pdf}
1335:    \hspace*{0.4cm}   
1336:    \includegraphics[width=8.7cm]{fig5b.pdf}
1337: }
1338: }
1339: \caption{
1340: The \RP{}-planes for the classic model (the left-hand panel) and for the
1341: generalized model (the right-hand panel) for the two-planet system analyzed in
1342: Fig.~\ref{fig:fig2}. The thick, red curves mark stationary modes in  the classic
1343: model, the thick blue curves are for the equilibria in the generalized model.
1344: Shaded areas indicate zones of $\Delta\varpi$ librations. The thick red lines
1345: are for the collision line of orbits. Colors code $\log$~SN of the {\em outer
1346: orbit}, characterizing  solutions derived numerically with the help of the
1347: direct code. Black points are for strongly chaotic solutions with the
1348: $\log\mbox{SN} \sim 3$, white color  is for regular solutions with
1349: $\log\mbox{SN}\sim 0$, intermediate values are marked with the color scale above
1350: the panels, accordingly (see the text for more details).
1351: }
1352: \label{fig:fig5}
1353: \end{figure*}
1354: \else
1355: \begin{figure*}
1356: \centerline{
1357: \hbox{\includegraphics[width=8.7cm]{fig5a.eps}
1358:    \hspace*{0.4cm}   
1359:    \includegraphics[width=8.7cm]{fig5b.eps}
1360: }
1361: }
1362: \caption{
1363: The \RP{}-planes for the classic model (the left-hand panel) and for the
1364: generalized model (the right-hand panel) for the two-planet system analyzed in
1365: Fig.~\ref{fig:fig2}. The thick, red curves mark stationary modes in  the classic
1366: model, the thick blue curves are for the equilibria in the generalized model.
1367: Shaded areas indicate zones of $\Delta\varpi$ librations. The thick red lines
1368: are for the collision line of orbits. Colors code $\log$~SN of the {\em outer
1369: orbit}, characterizing  solutions derived numerically with the help of the
1370: direct code. Black points are for strongly chaotic solutions with the
1371: $\log\mbox{SN} \sim 3$, white color  is for regular solutions with
1372: $\log\mbox{SN}\sim 0$, intermediate values are marked with the color scale above
1373: the panels, accordingly (see the text for more details).
1374: }
1375: \label{fig:fig5}
1376: \end{figure*} 
1377: \fi
1378: %
1379: \corr{
1380: Finally, to illustrate the development of the secular instability, we solved the
1381: equations of motion of the full system, starting very close to the UE lying on
1382: the energy level between levels {\bf e} and {\bf f}, as marked in
1383: Fig.~\ref{fig:fig2}, and Fig.~\ref{fig:fig4}e,f. The solution is illustrated in
1384: the phase diagram in the left-hand panel of Fig.~\ref{fig:fig6}. For a reference, the
1385: numerical solution is over-plotted on the analytically derived separatrices of
1386: the UE. The corresponding time evolution of the orbital elements are illustrated
1387: in the right-hand panels of Fig.~\ref{fig:fig6}. Clearly, during quite a long
1388: time the full system stays close to the UE, but after $\sim 10^5$~yrs it follows
1389: a trajectory  close to the inner separatrix, and finally begins to move close to
1390: the outer separatrix, approaching large $e_2$. During this evolution, we observe
1391: not only very irregular behavior of $a_2$ and both  eccentricities, but also
1392: $\Delta\varpi$ changing from large amplitude  librations around $0$ to
1393: circulations. Although the configuration seems bounded during many secular
1394: periods, such behavior may be classified as strongly chaotic.
1395: }
1396: 
1397: %
1398: \ifpdf
1399: \begin{figure*}
1400: \centerline{
1401: \vbox{
1402: \hbox{
1403: \includegraphics[height=61mm]{fig6a.png}\hskip5mm
1404: \includegraphics[height=61mm]{fig6b.png}
1405: }
1406: }
1407: }
1408: \caption{
1409: The phase diagram in the $(e_2\cos{\Delta{\varpi}}, e_2
1410: \sin{\Delta{\varpi}})$-plane (the left-hand panel) drawn to illustrate a development
1411: of the secularly unstable behavior of the full system.  The numerical solution
1412: of the full system (thin, black curve) is over-plotted on the analytical,
1413: secular solution (gray, thicker curve). The right-hand panels illustrate the
1414: relative changes of semi-major axes (top panel), eccentricities (middle panel)
1415: [grey curves are for outer orbit, black curves are for the inner orbit], and the
1416: apsidal angle $\Delta\varpi$ (bottom panel).  The initial orbital parameters are
1417: the same as in Fig. \ref{fig:fig2}.  The energy level of this solution lies
1418: between levels \textbf{e}-\textbf{f} marked in Fig.~\ref{fig:fig2}, accordingly,
1419: compare it also with phase diagrams Fig.~\ref{fig:fig4}e,f.
1420: }
1421: \label{fig:fig6}
1422: \end{figure*}
1423: \else
1424: \begin{figure*}
1425: \centerline{
1426: \vbox{
1427: \hbox{
1428: \includegraphics[height=61mm]{fig6a.eps}\hskip5mm
1429: \includegraphics[height=61mm]{fig6b.eps}
1430: }
1431: }
1432: }
1433: \caption{
1434: The phase diagram in the $(e_2\cos{\Delta{\varpi}}, e_2
1435: \sin{\Delta{\varpi}})$-plane (the left-hand panel) drawn to illustrate a development
1436: of the secularly unstable behavior of the full system.  The numerical solution
1437: of the full system (thin, black curve) is over-plotted on the analytical,
1438: secular solution (gray, thicker curve). The right-hand panels illustrate the
1439: relative changes of semi-major axes (top panel), eccentricities (middle panel)
1440: [grey curves are for outer orbit, black curves are for the inner orbit], and the
1441: apsidal angle $\Delta\varpi$ (bottom panel).  The initial orbital parameters are
1442: the same as in Fig. \ref{fig:fig2}.  The energy level of this solution lies
1443: between levels \textbf{e}-\textbf{f} marked in Fig.~\ref{fig:fig2}, accordingly,
1444: compare it also with phase diagrams Fig.~\ref{fig:fig4}e,f.
1445: }
1446: \label{fig:fig6}
1447: \end{figure*}
1448: \fi
1449: %
1450: %-------------------------------------------------------------------------------
1451: \section{Parametric survey of two-planet systems}
1452: %-------------------------------------------------------------------------------
1453: %
1454: The characterization of the phase space with the help of the representative
1455: plane  can be  very useful to  conduct a survey of the basic features of the
1456: secular dynamics. In particular, we want to understand how it depends on the 
1457: physical and orbital parameters governing the magnitude of the GR and QM
1458: interactions. In subsequent diagrams of the \RP{}-plane, we will always mark the
1459: collision and anti-collision lines. In this way, we can determine the border of
1460: validity of the analytic approach. Yet to derive the stationary modes possibly
1461: exactly, in the whole permitted range of eccentricity,  we compute their
1462: locations with the help of the semi-analytical averaging algorithm.
1463: %
1464: %-------------------------------------------------------------------------------
1465: \subsection{Dependence of the secular dynamics on the masses}
1466: \label{mass_dependence}
1467: %-------------------------------------------------------------------------------
1468: %
1469: The results of the survey of the secular dynamics of two-planet systems,
1470: including GR and QM interactions, for varied planetary masses, are illustrated
1471: in   Fig.~\ref{fig:fig7}. 
1472: %
1473: \ifpdf
1474: \begin{figure*}
1475: \centerline{
1476: \vbox{
1477: \hbox{\includegraphics[width=5.9cm]{fig7a.png}
1478:       \includegraphics[width=5.9cm]{fig7b.png}
1479:       \includegraphics[width=5.9cm]{fig7c.png}}
1480: \hbox{\includegraphics[width=5.9cm]{fig7d.png}
1481:       \includegraphics[width=5.9cm]{fig7e.png}
1482:       \includegraphics[width=5.9cm]{fig7f.png}}
1483: }
1484: }      
1485: \caption{
1486: The representative energy planes $(e_1 \cos{\Delta\varpi}, e_2)$ for
1487: $\Delta\varpi=0$ (the right half-planes,\RPp{}) and $\Delta\varpi=\pi$ (the left
1488: half-planes, \RPm{}). 
1489: Color contours are for the ratio of the  apsidal frequency induced by
1490: the general relativity and quadrupole moment  to the  apsidal frequency caused by mutual
1491: interactions between planets. 
1492: Thick curves mark positions of stationary solutions. Red
1493: lines are for stable equilibria in the classic model. Blue and violet 
1494: curves mark
1495: the positions of stable and unstable equilibria of the generalized model. The
1496: thin lines are for the secular energy levels of the generalized model. The thick, skew lines are for the
1497: collision lines defined with $a_1 (1 \pm e_1) = a_2 (1 - e_2)$. Parameters of
1498: these systems are as follows: $m_0 = 1 \mbox{M}_{\sun}$, $a_1=0.1 \mbox{au}$,
1499: $a_2=1 \mbox{au}$, $R_0=1 \mbox{R}_{\sun}$,  $T_{\idm{rot}}=30$~days,
1500: $k_L=0.02$.  Each panel was calculated for varied planetary masses, under the
1501: condition of constant mass ratio $m_1/m_2 = 3$. The mass of the inner planet is
1502: written in the top left corner of each panel.
1503: }
1504: \label{fig:fig7}
1505: \end{figure*}
1506: \else
1507: \begin{figure*}
1508: \centerline{
1509: \vbox{
1510: \hbox{\includegraphics[width=5.9cm]{fig7a.eps}
1511:       \includegraphics[width=5.9cm]{fig7b.eps}
1512:       \includegraphics[width=5.9cm]{fig7c.eps}}
1513: \hbox{\includegraphics[width=5.9cm]{fig7d.eps}
1514:       \includegraphics[width=5.9cm]{fig7e.eps}
1515:       \includegraphics[width=5.9cm]{fig7f.eps}}
1516: }
1517: }      
1518: \caption{
1519: The representative energy planes $(e_1 \cos{\Delta\varpi}, e_2)$ for
1520: $\Delta\varpi=0$ (the right half-planes,\RPp{}) and $\Delta\varpi=\pi$ (the left
1521: half-planes, \RPm{}). 
1522: Color contours are for the ratio of the  apsidal frequency induced by
1523: the general relativity and quadrupole moment  to the  apsidal frequency caused by mutual
1524: interactions between planets. 
1525: Thick curves mark positions of stationary solutions. Red
1526: lines are for stable equilibria in the classic model. Blue and violet 
1527: curves mark
1528: the positions of stable and unstable equilibria of the generalized model. The
1529: thin lines are for the secular energy levels of the generalized model. The thick, skew lines are for the
1530: collision lines defined with $a_1 (1 \pm e_1) = a_2 (1 - e_2)$. Parameters of
1531: these systems are as follows: $m_0 = 1 \mbox{M}_{\sun}$, $a_1=0.1 \mbox{au}$,
1532: $a_2=1 \mbox{au}$, $R_0=1 \mbox{R}_{\sun}$,  $T_{\idm{rot}}=30$~days,
1533: $k_L=0.02$.  Each panel was calculated for varied planetary masses, under the
1534: condition of constant mass ratio $m_1/m_2 = 3$. The mass of the inner planet is
1535: written in the top left corner of each panel.
1536: }
1537: \label{fig:fig7}
1538: \end{figure*}
1539: \fi
1540: %
1541: We fix the system parameters as follows: the mass of the  parent star is $m_0 =
1542: 1~\mbox{M}_{\sun}$,  the equatorial radius of the star $R_0 = 
1543: 1~\mbox{R}_{\sun}$, and  $T_{\idm{rot}}=30$~days, $k_L=0.02$ (then $J_2 \sim
1544: 10^{-7}$),  the semi-major axes of the planets are $a_1 = 0.1~\mbox{au}$,
1545: and $a_2 = 1.0~\mbox{au}$, respectively. Hence, we consider a typical
1546: hierarchical configuration with the orbital periods ratio $\sim 30$. In this
1547: experiment, the planetary masses are varied, but their ratio is kept constant,
1548: $m_1/m_2=3$. For a reference, the mass of the inner planet is labeled in the
1549: top-left corner of each respective panel in Fig.~\ref{fig:fig7}. Basically, in
1550: this test we can also analyze the effect of unknown inclination of the co-planar
1551: system on the long-term dynamics and stability. However, as we show in a recent
1552: work regarding the 14~Herculis planetary system \citep{Gozdziewski2008},
1553: planetary masses  derived from observations do not necessarily always scale
1554: according to the law of the mass-factor $1/\sin~i$. If the minimal masses are
1555: large then the mutual interactions in low-inclination configurations can
1556: strongly modify the RV signal and even {\em the mass hierarchy} may be reversed 
1557: in the orbital fits.
1558: 
1559: Figure~\ref{fig:fig7} reveals that curves representing stationary modes,  which
1560: are known in the classic problem, are usually significantly shifted and/or
1561: distorted.  Also new features of the \RP{}-plane appear and  it can be seen in
1562: the top-left panel of Fig.~\ref{fig:fig7}. In general,  the distortions of
1563: equilibria curves are the more stronger when the masses are smaller. It is quite
1564: straightforward to explain this effect. When the planetary masses decrease, also
1565: their mutual interactions (scaled by $m_1 m_2$) are decreasing.  Yet the
1566: pericenter frequency induced by $\left<\Hmut\right>$ is scaled by the mass
1567: product. Simultaneously, the GR and the spin-induced apsidal frequencies do not depend on the
1568: planetary masses directly (see Eq.~\ref{orel} and  Eq.~\ref{oqm}, respectively),
1569: and they can be regarded as approximately constant in the given mass range.
1570: Therefore  $\kappa$ increases with decreasing $m_1$ and $m_2$. 
1571: \corr{Then, also the assumptions of the secular theory are better
1572: fulfilled.}
1573: 
1574: Some parts of the stationary curves in the \RPm{} half-plane (for
1575: $\Delta{\varpi}=\pi$) comprise of unstable equilibria (they are marked with
1576: violet color). As we mentioned already, to the best of our knowledge, such
1577: solutions are yet unknown in the literature. Similarly to the non-classic
1578: equilibria discovered by \cite{Michtchenko2004}, these solutions are accompanied
1579: by the \TSR{} solutions and correspond to saddles of the secular
1580: Hamiltonian.
1581: The behavior of neighboring solutions tells us that they are Lyapunov unstable.
1582: This has been analyzed  in Sect.~3.
1583: 
1584: 
1585: Actually, the sequence of panels in Fig.~\ref{fig:fig7} illustrates a
1586: characteristic development of curves representing the  equilibria, including the
1587: \UE{} solutions. When the masses are relatively large (see the top-left panel of
1588: Fig.~\ref{fig:fig7}), the equilibria  curves are distorted and the unstable
1589: equilibria appear at the very edge of the \RPm{}-plane, in the range of moderate
1590: and large values of eccentricity. On contrary, in the classic model, the
1591: \UE{} solutions
1592: can appear (in fact, they were found) only for $\Delta\varpi=0$
1593: \citep{Michtchenko2004}.
1594: %
1595: %%In the analyzed range of masses, it does not exist in the generalized model. 
1596: %
1597: Seemingly, the new \UE{} branch located in the \RPm{}-plane is
1598: specific only for this model. When the masses decrease then $\kappa$ grows (so
1599: the GR+QM effects become comparable in magnitude to the NG interactions). This
1600: leads to further distortion of mode~II curves and to expanding the \UE{} part
1601: towards moderate $e_1$. At some point (between $m_1\sim 1.8~\mbox{m}_{\idm{J}}$
1602: and $m_1\sim 1.2~\mbox{m}_{\idm{J}}$) both  stationary curves meet in a
1603: bifurcation point. Here, we can explain the particular choice of parameters used
1604: to construct  Fig.~\ref{fig:fig2}. When the masses become smaller, the 
1605: equilibria curves separate along $e_2$. We note that already for  $m_1\sim
1606: 1.2~\mbox{m}_{\idm{J}}$,  the \RP{}-plane is dominated by the GR+QM corrections.
1607: We recall that in the classic case, the qualitative features of the \RP{}-plane
1608: do not depend on the masses individually \citep{Michtchenko2004}, only on their
1609: ratio in the approximation of small values (see also the sequence of plots in
1610: Fig.~\ref{fig:fig7}). This conclusion is not true
1611: anymore in the
1612: realm of the generalized model.
1613: %
1614: %-----------------------------------------------------------------------------
1615: \subsection{Dependence of the secular dynamics on semi-major axes}
1616: %-----------------------------------------------------------------------------
1617: %
1618: In the next experiment, we investigate the dependence of the  secular dynamics
1619: of the generalized model on individual semi-major axes;
1620: note that the dynamics of the classic model depend only
1621: on their ratio, $\alpha$. The results are illustrated in
1622: Fig.~\ref{fig:fig8}. We proceed in the same manner as to draw 
1623: Figs.~\ref{fig:fig2} and \ref{fig:fig7}. We seek for stationary solutions, and we
1624: overplot the found equilibria on color-coded  contour levels of coefficient
1625: $\kappa$. The primary parameters of the tested configurations are the
1626: following:  $m_0 = 1~\mbox{M}_{\sun}$, $R = 1~\mbox{R}_{\sun}$,
1627: $T_{\idm{rot}}=30$~days, $k_L=0.02$, $m_1 = 0.4~\mbox{m}_{\idm{J}}$,  $m_2 =
1628: 0.2~\mbox{m}_{\idm{J}}$.  The ratio of semi-major axes is kept constant, $\alpha
1629: \equiv a_1/a_2 = 0.1$, while the individual $a_1, a_2$ are varied.   For a
1630: reference, the nominal value of $a_1$ is labeled in the top-left corner in each
1631: respective panel.  
1632: %
1633: \ifpdf
1634: \begin{figure*}
1635: \centerline{
1636: \vbox{
1637: \hbox{\includegraphics[width=5.9cm]{fig8a.png}
1638:       \includegraphics[width=5.9cm]{fig8b.png}
1639:       \includegraphics[width=5.9cm]{fig8c.png}}
1640: \hbox{\includegraphics[width=5.9cm]{fig8d.png}
1641:       \includegraphics[width=5.9cm]{fig8e.png}
1642:       \includegraphics[width=5.9cm]{fig8f.png}}
1643:       }
1644:       }
1645: \caption{
1646: The survey of the secular dynamics of two-planet system, when the semi-major
1647: axes are varied. Panels are constructed in the same way,  as in Fig.
1648: \ref{fig:fig7}.  
1649: Color contours are for the ratio of the  apsidal frequency induced by
1650: the general relativity and quadrupole moment  to the  apsidal frequency caused by mutual
1651: interactions between planets. 
1652: The thick skew
1653: line is for collisions line defined with $a_1 (1 \pm e_1) = a_2 (1 - e_2)$. The
1654: masses are $m_0 = 1 \mbox{M}_{\sun}$, $m_1=0.4 \mbox{m}_{\idm{J}}$, $m_2=0.2
1655: \mbox{m}_{\idm{J}}$, respectively, the characteristic radius of the star  $R_0=1
1656: \mbox{R}_{\sun}$, $T_{\idm{rot}}=30$~days, $k_L=0.02$. Each panel is for
1657: different semi-major axes, fulfilling the condition of constant $\alpha \equiv
1658: a_1/a_2 = 0.1$.  The semi-major axis of the inner planet labels each respective
1659: panel. Compared to  Fig. \ref{fig:fig7}, an additional  unstable equilibrium for
1660: the classic model appears and it is marked with thick green lines.
1661: }
1662: \label{fig:fig8}
1663: \end{figure*}
1664: \else
1665: \begin{figure*}
1666: \centerline{
1667: \vbox{
1668: \hbox{\includegraphics[width=5.9cm]{fig8a.eps}
1669:       \includegraphics[width=5.9cm]{fig8b.eps}
1670:       \includegraphics[width=5.9cm]{fig8c.eps}}
1671: \hbox{\includegraphics[width=5.9cm]{fig8d.eps}
1672:       \includegraphics[width=5.9cm]{fig8e.eps}
1673:       \includegraphics[width=5.9cm]{fig8f.eps}}
1674:       }
1675:       }
1676: \caption{
1677: The survey of the secular dynamics of two-planet system, when the semi-major
1678: axes are varied. Panels are constructed in the same way,  as in Fig.
1679: \ref{fig:fig7}.  
1680: Color contours are for the ratio of the  apsidal frequency induced by
1681: the general relativity and quadrupole moment  to the  apsidal frequency caused by mutual
1682: interactions between planets. 
1683: The thick skew
1684: line is for collisions line defined with $a_1 (1 \pm e_1) = a_2 (1 - e_2)$. The
1685: masses are $m_0 = 1 \mbox{M}_{\sun}$, $m_1=0.4 \mbox{m}_{\idm{J}}$, $m_2=0.2
1686: \mbox{m}_{\idm{J}}$, respectively, the characteristic radius of the star  $R_0=1
1687: \mbox{R}_{\sun}$, $T_{\idm{rot}}=30$~days, $k_L=0.02$. Each panel is for
1688: different semi-major axes, fulfilling the condition of constant $\alpha \equiv
1689: a_1/a_2 = 0.1$.  The semi-major axis of the inner planet labels each respective
1690: panel. Compared to  Fig. \ref{fig:fig7}, an additional  unstable equilibrium for
1691: the classic model appears and it is marked with thick green lines.
1692: }
1693: \label{fig:fig8}
1694: \end{figure*}
1695: \fi
1696: %
1697: For decreasing $a_1$, the derivatives of $\Hmut, \Hrel, \Hflat$ over $G_1$ 
1698: increase, hence the magnitude of the respective correction to 
1699: the apsidal  frequency grows.
1700: Moreover, the GR and QM induced apsidal frequency increase faster than the rate
1701: of the pericenter advance forced by the NG interactions. In the region of
1702: \RP{}-plane
1703: painted in yellow (light gray), the GR and QM perturbations dominate over the NG
1704: interactions.  This region expands quickly with decreasing semi-major axis of
1705: the inner planet, $a_1$. Already for $a_1 \sim 0.4$~au (which is similar to the
1706: semi-major axis of Mercury in the Solar system), the apparent {\em corrections}
1707: to the classic model, may contribute much more to the pericenter frequency  of
1708: the innermost planet than the point mass Newtonian interactions.
1709: 
1710: The top-left panel in Fig. \ref{fig:fig8} is for  $a_1 = 2$~au and $a_2 =
1711: 20$~au, respectively. For these parameters, a shift of the curve of stationary
1712: solutions, when compared to the ones in the classic model, is already
1713: significant. In the next  panel ($a_1=1$~au, $a_2=10$~au) the distortion of
1714: curves representing stationary modes is even stronger. Moreover, the \UE{} mode
1715: present in the classic model in the \RPp{}-plane, cannot be found  in that
1716: half-plane anymore. Simultaneously, new solutions appear at the very edge of
1717: the  \RPm-plane, in the range of large $e_1$. For $a_1=0.5$~au, this branch of
1718: stationary modes is even more extended. Starting with this value of  $a_1$, the
1719: structure of the \RP{}-plane with respect to the generalized  model is very
1720: different from that ones in the classic case. For smaller $a_1$, the curves of 
1721: stationary modes are still more distorted.  Clearly, these distortions cannot be
1722: regarded as small. This result is  quite unexpected, recalling that the
1723: semi-major axes and the planetary masses by no means  are ``extreme''. In spite
1724: of these ``typical'' parameters, the secular theories of the classic and
1725: generalized models lead to qualitatively different view of the phase space. We
1726: stress again that the notion of the GR and QM effects as corrections (or small
1727: perturbations) to the  secular Hamiltonian should be understood in quite a new
1728: light.
1729: %
1730: %-------------------------------------------------------------------------------
1731: \subsection{Dependence of the secular dynamics on the stellar spin}
1732: %-------------------------------------------------------------------------------
1733: %
1734: In the last parametric survey,  we study the dependence of the secular dynamics
1735: in the realm of the generalized model on the stellar spin (or, effectively, on
1736: the second zonal harmonic $J_2$). Figure \ref{fig:fig9} illustrates the
1737: \RP{}-plane  computed for  following  parameters: $m_0 = 1~\mbox{M}_{\sun}$, \
1738: $m_1=1.25~\mbox{m}_{\idm{J}}$, $m_2=0.25~\mbox{m}_{\idm{J}}$, $a_1=0.1$~au,
1739: $a_2=1.0$~au,  $R_0=1~\mbox{R}_{\sun}$, $k_L=0.02$. The top-left panel in
1740: Fig.~\ref{fig:fig9} is for the GR correction only and subsequent plots are for
1741: decreasing rotation period (generalized model)  $T_{\idm{rot}}$ of the star (its particular values 
1742: label the respective plots). This sequence corresponds to increasing $J_2$.
1743: %
1744: \ifpdf
1745: \begin{figure*}
1746: \centerline{
1747: \vbox{
1748: \hbox{\includegraphics[width=5.9cm]{fig9a.png}
1749:       \includegraphics[width=5.9cm]{fig9b.png}
1750:       \includegraphics[width=5.9cm]{fig9c.png}
1751:       }
1752: \hbox{\includegraphics[width=5.9cm]{fig9d.png}
1753:       \includegraphics[width=5.9cm]{fig9e.png}
1754:       \includegraphics[width=5.9cm]{fig9f.png}
1755:       }
1756: }}      
1757: \caption{
1758: The representative energy planes $(e_1 \cos{\Delta\varpi}, e_2)$ for 
1759: $\Delta\varpi=0$ (the right half-planes, \RPp{}) and $\Delta\varpi=\pi$  (the
1760: left half-planes, \RPm{}). Colors are for the ratio of the  apsidal frequency
1761: induced by the general relativity and quadrupole moment  to the  apsidal
1762: frequency caused by mutual interactions between planets.  The thick lines mark
1763: the stationary solutions. The thick red curves are for stable equilibria in the
1764: classic model. The blue and violet curves mark  positions of stable and unstable
1765: equilibria in the generalized model, respectively.  Thin lines are for the energy levels of a
1766: generalized model. The thick skew line is for collisions line defined with $a_1
1767: (1 \pm e_1) = a_2 (1 - e_2)$. Parameters of the system are the following: $m_0 =
1768: 1.0~\mbox{M}_{\sun}$,  $m_1=1.25~\mbox{m}_{\idm{J}}$,
1769: $m_2=0.25~\mbox{m}_{\idm{J}}$,  $a_1=0.1 \mbox{au}$, $a_2=1.0~\mbox{au}$, 
1770: $R_0=1~\mbox{R}_{\sun}$, $k_L=0.02$.  Subsequent panels are for  different
1771: rotational periods of the star, $T_{\idm{rot}}$ which is labeled in each
1772: respective plot. 
1773: \corr{For a reference,
1774: the top left panel is for the GR correction only.}
1775: }
1776: \label{fig:fig9}
1777: \end{figure*}
1778: \else
1779: \begin{figure*}
1780: \centerline{
1781: \vbox{
1782: \hbox{\includegraphics[width=5.9cm]{fig9a.eps}
1783:       \includegraphics[width=5.9cm]{fig9b.eps}
1784:       \includegraphics[width=5.9cm]{fig9c.eps}
1785:       }
1786: \hbox{\includegraphics[width=5.9cm]{fig9d.eps}
1787:       \includegraphics[width=5.9cm]{fig9e.eps}
1788:       \includegraphics[width=5.9cm]{fig9f.eps}
1789:       }
1790: }}      
1791: \caption{
1792: The representative energy planes $(e_1 \cos{\Delta\varpi}, e_2)$ for 
1793: $\Delta\varpi=0$ (the right half-planes, \RPp{}) and $\Delta\varpi=\pi$  (the
1794: left half-planes, \RPm{}). Colors are for the ratio of the  apsidal frequency
1795: induced by the general relativity and quadrupole moment  to the  apsidal
1796: frequency caused by mutual interactions between planets.  The thick lines mark
1797: the stationary solutions. The thick red curves are for stable equilibria in the
1798: classic model. The blue and violet curves mark  positions of stable and unstable
1799: equilibria in the generalized model, respectively.  Thin lines are for the energy levels of a
1800: generalized model. The thick skew line is for collisions line defined with $a_1
1801: (1 \pm e_1) = a_2 (1 - e_2)$. Parameters of the system are the following: $m_0 =
1802: 1.0~\mbox{M}_{\sun}$,  $m_1=1.25~\mbox{m}_{\idm{J}}$,
1803: $m_2=0.25~\mbox{m}_{\idm{J}}$,  $a_1=0.1 \mbox{au}$, $a_2=1.0~\mbox{au}$, 
1804: $R_0=1~\mbox{R}_{\sun}$, $k_L=0.02$.  Subsequent panels are for  different
1805: rotational periods of the star, $T_{\idm{rot}}$ which is labeled in each
1806: respective plot. 
1807: \corr{For a reference,
1808: the top left panel is for the GR correction only.}
1809: }
1810: \label{fig:fig9}
1811: \end{figure*}
1812: \fi
1813: %
1814: The top-left panel of Fig.~\ref{fig:fig9} is for a spherical,
1815: non-rotating star ($J_2=0$, no tidal bulge),  the bottom-right panel is for 
1816: very fast rotation characteristic for a young object (then
1817: $J_2=1\times10^{-4}$). Curiously,  changes of the spin encompassing that range
1818: are well enough to emerge new families and  stationary solutions which we
1819: described above. They are represented, as before, by thick violet curves drawn in the
1820: \RPm{}-plane.  Actually,  the sequence of plots simulates  variations of
1821: flattening during the lifetime of the star. Hence, after skipping 
1822: dissipative tidal perturbations, we may conclude that the structure of the phase space of
1823: the secular system (and, in general, also its dynamical stability) may depend
1824: not only on the observed or measured  orbital parameters but also on the age and
1825: the spin rate of the host star. 
1826: 
1827: The structure of the \RP{}-plane and stationary modes which are illustrated
1828: in Fig.~\ref{fig:fig9} remind us closely Fig.~\ref{fig:fig7} and
1829: Fig.~\ref{fig:fig8}. In fact, as we already mentioned,  the dependence of the dynamics on the model
1830: parameters may be described uniformly through  $\kappa$, see Eq.~\ref{eq:kappa}.
1831: Indeed, the increase of the stellar spin leads to increase of the  pericenter
1832: frequency and the nominator of Eq.~\ref{eq:kappa} (note that other terms are
1833: constant). The same behavior of $\kappa$ is caused by decreasing
1834: masses (see the sequence of diagrams in Fig.~\ref{fig:fig7} and also
1835: discussion in Sect.~4.1). In that
1836: case, two terms of the nominator of $\kappa$ do not change, but the denominator
1837: decreases. Finally, if the semi-major axes decrease in constant ratio, $\kappa$
1838: also grows because the GR and QM-induced  correction to the apsidal frequency
1839: grows {\em faster} than the NG-forced apsidal frequency. In that sense, all 
1840: point-mass gravity corrections governed by the described above parameter 
1841: changes, modify similarly the structure of the \RP-plane. 
1842: 
1843: Considering our simplified secular model, the existence of the branch
1844: of stationary solutions in
1845: the regime of large $e_1$ and small $e_2$ may seem questionable in {\em the real}
1846: planetary configurations. In that regime, the tidal perturbations
1847: may become very significant for the secular evolution.  However,
1848: the position of the bifurcation point 
1849: in the \RPm{}-plane, where the two branches
1850: of equilibria curves meet, and which marks the extent of the branch, is
1851: shifted towards moderate $e_1 \sim 0.6$--$0.7$ when the  mass
1852: ratio $m_1/m_2$  grows.
1853: This effect can be observed in the geometric evolution of
1854: these branches in Figs.~\ref{fig:fig8}, \ref{fig:fig7} and \ref{fig:fig9}
1855: which form a sequence of $m_1/m_2=2,3,5$, respectively.
1856: Hence, planetary configurations in that regime may really be found  
1857: in the Nature.
1858: 
1859: %
1860: %-------------------------------------------------------------------------------
1861: \section{The secular dynamics of the $\upsilon$~And system}
1862: %-------------------------------------------------------------------------------
1863: %
1864: Finally, we consider the generalized coplanar problem of three planets.  This
1865: model can be still described by the Hamiltonian written  in the general form of
1866: Eqs.~(\ref{HKepl}--\ref{Hrel}), and (\ref{eq:hrf}--\ref{eq:htb}) with $N=3$. The
1867: averaged perturbing Hamiltonian has the form of Eqs.~(\ref{Hsec}),
1868: (\ref{avHmut}), (\ref{expanssion}), (\ref{avbHrel}), (\ref{avHflat}) and
1869: (\ref{avHflatT}). To study the properties of the secular system, we introduce  
1870: the following set of angle-action variables related to the Poincar\'e canonical
1871: elements
1872: \citep{Migaszewski2008a}:
1873: \begin{eqnarray}
1874: \label{3pl}
1875: && \sigma_1 \equiv \varpi_3 - \varpi_1,  \quad G'_1, \nonumber\\
1876: && \sigma_2 \equiv \varpi_3 - \varpi_2 ,  \quad G'_2, \\
1877: && \sigma_3 \equiv -\varpi_3,  \quad \mbox{AMD} = G'_1 + G'_2 + G'_3, \nonumber
1878: \end{eqnarray}
1879: where $G'_i \equiv L_i-G_i$ (see Eq.~\ref{dvars}). Because $\sigma_3$ is the
1880: cyclic angle, the Angular Momentum Deficit (AMD) is conserved, and the reduced
1881: system has two degrees of freedom. These variables make it possible to construct
1882: the representative energy plane in a similar manner as  for the two-planet
1883: system. Moreover, the choice of such a plane is not unique. One of possible
1884: definitions can be  derived through a direct analogy to the two-planet system
1885: case. The {\em symmetric} representative plane can be defined as the set of
1886: phase-space  points fulfilling the following relation:
1887: \[
1888: \frac{\partial\,\Hsec}{\partial\,\sigma_i} = 0, \quad i=1,2,3,
1889: \]
1890: with the simultaneous conditions that $\sigma_i=0,\pi$. For details, see
1891: \citep{Migaszewski2008a}.
1892: 
1893: A sequence of symmetric  representative planes shown in Fig.~\ref{fig:fig10}
1894: illustrates the qualitative properties of the generalized secular model of the
1895: $\upsilon$ Andromedae planetary system \citep{Butler1999}. This system comprises
1896: of three planets with masses and semi-major axes  derived through the radial
1897: velocity observations:
1898: $m_0=1.27~\mbox{M}_{\sun}$,
1899: $m_1=0.69~\mbox{m}_{\idm{J}}$, 
1900: $m_2=1.98~\mbox{m}_{\idm{J}}$, 
1901: $m_3=3.95~\mbox{m}_{\idm{J}}$, 
1902: $a_1=0.059~\mbox{au}$, 
1903: $a_2=0.83~\mbox{au}$, 
1904: $a_3=2.51~\mbox{au}$.
1905: The constant of the AMD integral (effectively, the integral
1906: of the total angular momentum) was obtained for the following eccentricities of the
1907: nominal system:  $e_1=0.029$, $e_2=0.254$, $e_3=0.242$. 
1908: %For a reference, the top-left panel in Fig.~\ref{fig:fig10} is computed for the classic model.
1909: 
1910: We consider the representative plane for varying age of the parent star,
1911: starting with approximately 30~Myr before entering the ZAMS.  Subsequent plots
1912: are labeled by the lifetime $\tau$ relative to this moment taken  as the
1913: zero-point of the time scale, rotation period $T_{\idm{rot}}$ and stellar radius
1914: ($R_{0}$) expressed in Sun's radius. For a reference, the  classic model and
1915: with the GR correction  are illustrated in the first two top-left panels of
1916: Fig.~\ref{fig:fig10}. These panels are derived for  non-rotating, non-distorted
1917: spherical star. When the star is spinning, its second zonal harmonic may be as
1918: large $J_2 \sim  10^{-3}$.  The current equatorial  radius 
1919: of $\upsilon$~Andr is approximately 
1920: $R_0=1.26~\mbox{R}_{\sun}$. We note that the stellar radius and the spin period  of
1921: the star at $\tau=-30$~Myr are taken from \citep{Nagasawa2005}, and were
1922: linearly interpolated over $\tau \in [-30,0]$~Myr.
1923: 
1924: The results are again quite surprising.  After adding the GR corrections to the
1925: Hamiltonian of the classic model, the  overall view of the phase space changes
1926: significantly.  The saddle of the secular Hamiltonian which is present  in the
1927: classic model now vanishes.  At its place, a new {\em maximum} of the secular
1928: Hamiltonian appears. Moreover, at the bottom half-plane of the representative
1929: plane, close to the border of the permitted region of motion,  two new saddle
1930: points appear. 
1931: 
1932: The next diagrams illustrate changes of the structure of the \RP{}-plane and a
1933: development  of equilibria in a sequence simulating time evolution of the
1934: stellar spin.  At the beginning, before the star enters the ZAMS, the
1935: characteristic plane reveals a sharp maximum and a saddle in the very edge of
1936: the region of permitted motions.  The thin curve surrounding the maximum marks
1937: the energy level of the nominal system. When the rotation period increases
1938: up to $\sim 8$~days, the secondary extremum (the minimum) emerges in the place of the
1939: saddle and it persists shortly before the ZAMS stage and for longer rotational
1940: periods.
1941: 
1942: Curiously, the only feature seen in the energy diagrams, which survives the spin
1943: variations during the whole lifetime of the star, and persist in  the
1944: generalized model  (with the GR and QM corrections) is the stable equilibrium
1945: point related to the maximum of the secular Hamiltonian, which  is found in the
1946: range of  small $e_1$ and moderate $e_2$. Curiously, the nominal system appears
1947: in the energy level drawn with grey (green), thick line surrounding this
1948: maximum of $\Hsec$. 
1949: \corr{We also may notice that close to
1950: this equilibrium of the generalized model, a
1951: saddle of the classic model appears which is {\em
1952: linearly stable}. 
1953: }
1954: %
1955: \ifpdf
1956: \begin{figure*}
1957: \centerline{
1958: \vbox{
1959: \hbox{
1960: \includegraphics[width=44mm]{fig10a.pdf}
1961: \includegraphics[width=44mm]{fig10b.pdf}
1962: \includegraphics[width=44mm]{fig10c.pdf}
1963: \includegraphics[width=44mm]{fig10d.pdf}
1964: }
1965: \hbox{
1966: \includegraphics[width=44mm]{fig10e.pdf}
1967: \includegraphics[width=44mm]{fig10f.pdf}
1968: \includegraphics[width=44mm]{fig10g.pdf}
1969: \includegraphics[width=44mm]{fig10h.pdf}
1970: }
1971: }}
1972: \caption{
1973: The secular energy levels on the \textit{symmetric representative plane} for
1974: three-planet system $\upsilon$ Andromede. The map coordinates are $x \equiv e_1
1975: \cos{\sigma_1}$, $y \equiv e_2 \cos{\sigma_2}$, where $ \sigma_1, \sigma_2$ are
1976: $0$ (positive values of $x$ or $y$) or $\pi$ (negative values of $x$ or $y$).
1977: Grey-colored region means forbidden motions with 
1978: $e_3<0$. Black curves are for the energy levels, green, thick
1979: levels are for the energy of the nominal $\upsilon$~Andr  system. Each panel is for a different
1980: setup of the planetary system model.  
1981: From the {\em left-top panel}: the first panel is for
1982: the classic NG model, the next panel is for the NG+GR model. 
1983: Next panels are for generalized model with the QM corrections parameterized
1984: by the spin rate of the parent star and its lifetime $\tau$ before
1985: the ZAMS stage. 
1986: }
1987: \label{fig:fig10}
1988: \end{figure*}
1989: \else
1990: \begin{figure*}
1991: \centerline{
1992: \vbox{
1993: \hbox{
1994: \includegraphics[width=44mm]{fig10a.eps}
1995: \includegraphics[width=44mm]{fig10b.eps}
1996: \includegraphics[width=44mm]{fig10c.eps}
1997: \includegraphics[width=44mm]{fig10d.eps}
1998: }
1999: \hbox{
2000: \includegraphics[width=44mm]{fig10e.eps}
2001: \includegraphics[width=44mm]{fig10f.eps}
2002: \includegraphics[width=44mm]{fig10g.eps}
2003: \includegraphics[width=44mm]{fig10h.eps}
2004: }
2005: }}
2006: \caption{
2007: The secular energy levels on the \textit{symmetric representative plane} for
2008: three-planet system $\upsilon$ Andromede. The map coordinates are $x \equiv e_1
2009: \cos{\sigma_1}$, $y \equiv e_2 \cos{\sigma_2}$, where $ \sigma_1, \sigma_2$ are
2010: $0$ (positive values of $x$ or $y$) or $\pi$ (negative values of $x$ or $y$).
2011: Grey-colored region means forbidden motions with 
2012: $e_3<0$. Black curves are for the energy levels, green, thick
2013: levels are for the energy of the nominal $\upsilon$~Andr  system. Each panel is for a different
2014: setup of the planetary system model.  
2015: From the {\em left-top panel}: the first panel is for
2016: the classic NG model, the next panel is for the NG+GR model. 
2017: Next panels are for generalized model with the QM corrections parameterized
2018: by the spin rate of the parent star and its lifetime $\tau$ before
2019: the ZAMS stage. 
2020: }
2021: \label{fig:fig10}
2022: \end{figure*}
2023: \fi
2024: %
2025: %-------------------------------------------------------------------------------
2026: \section{Conclusions}
2027: %-------------------------------------------------------------------------------
2028: %
2029: In this work, we consider a generalized secular theory of coplanar, $N$-planet
2030: system. Extending the model analyzed in the recent works devoted 
2031: to the secular planetary dynamics  with mutual Newtonian
2032: point-to-point interactions
2033: \citep[e.g.,][]{Michtchenko2004,Libert2005,Gallardo2005,Migaszewski2008a},  we
2034: consider the influence of the general relativity and quadrupole moment of the
2035: parent star on the secular dynamics of the innermost planet and stability of 
2036: the whole planetary system. In general, these corrections to the classic model
2037: still do not cover all physics governing the dynamics of such systems. In some
2038: cases (for instance, of the short-period hot-Jupiters), the tidal, dissipative
2039: torques acting between the inner planet and the star may be significant for the
2040: orbital evolution. However,  our main  goal is rather to extend the classic
2041: model with the perturbations that are conservative and may be well modeled in
2042: the realm of the Hamiltonian mechanics than to build a complete, general
2043: secular theory. Still, this approach is useful to a wide class of systems, when
2044: the tidal interactions may be regarded as secondary effects, or are acting 
2045: during much longer characteristic time-scale than the GR and QM perturbations.
2046: In reward, for paying the price of less general model, we may investigate the
2047: secular dynamics in a global manner. 
2048: 
2049: Our analytic model follows assumptions required by the averaging theorem.
2050: Technically, the averaging has been done with  the help of a very simple method.
2051: This algorithm relies on  appropriate change of integration variables. It does
2052: not incorporate any classic Fourier expansion of the perturbing function.  We
2053: obtain a very precise analytic model of the coplanar, $N$-planet system in
2054: terms of the semi-major axes ratio. It can be regarded as a generalization of
2055: the recent analytic secular theories  of the classic model investigated in
2056: many recent papers \citep[e.g.,][]{Ford2000,Lee2003,Michtchenko2004,Libert2005}. On the other
2057: hand,  our work also covers, to some extent, the global dynamics  of the
2058: generalized model studied in \cite{Mardling2002,Nagasawa2005} with the help of
2059: the Gauss/Lagrange planetary equations of motion. We stress, however, that our
2060: investigations are devoted to  more narrow class of systems (regarding the
2061: conservative perturbations).
2062: 
2063: A general conclusion which can be derived on the basis of the generalized theory
2064: is quite unexpected. Even in a case when the orbital parameters cannot be
2065: regarded as {\em extreme}, the corrections to the classic Hamiltonian stemming
2066: from the general relativity and the quadrupole moment of the star, may affect
2067: the secular dynamics dramatically. Not only  the structure of the phase space of
2068: the secular model changes, and new branches  of stationary solutions appear.
2069: These solutions may bifurcate within small relative ranges of the parameters
2070: (for instance, when the spin of the parent star is changing). We show that there
2071: is no simple and general recipe to predict the behavior of the secular system,
2072: when the perturbations are ``switched on''. The secular dynamics of  the
2073: generalized model becomes extremely complex and rich. For some combinations of
2074: the system parameters, the notion of the GR and QM  effects as {\em
2075: corrections}  to the point-mass  NG interactions does not seem proper anymore.  In some
2076: cases, these effects may be {\em more important} for the secular dynamics than
2077: the mutual, point mass Newtonian interactions between the planets. 
2078: 
2079: We also show that these effects may be significant for the dynamical stability
2080: of planetary systems both in the short-term and 
2081: in the secular time scales. For instance,
2082: the QM generated perturbations may directly depend on the star age and its
2083: physical parameters ($k_L$, $R_0$). In turn, these effects may strongly
2084: influence the structure of the phase space and can imply short-term, strong
2085: chaotic orbital evolution during a few secular periods.
2086: 
2087: The direct tests of the analytic theory are very encouraging.  The results
2088: justify its great accuracy. The precision of the analytic calculations is  very
2089: important for studying the global dynamics of hierarchical systems. The
2090: alternative numerical approach would require huge CPU time because the
2091: hierarchical planetary systems evolve during very different time scales. Then
2092: the CPU requirements of the direct numerical integrations are  by orders of
2093: magnitude larger than those ones needed by the analytic formulae. 
2094: %
2095: %-------------------------------------------------------------------------------
2096: \section*{Acknowledgments}
2097: %-------------------------------------------------------------------------------
2098: %
2099: We are very grateful to Rosemary Mardling for careful reading of the manuscript,
2100: constructive and informative review, many suggestions and invaluable remarks
2101: that greatly improved the work. We would like to thank Tatiana Michchenko for a
2102: discussion and comments on the manuscript. This work is supported by the Polish
2103: Ministry of Science and Education, Grant No. 1P03D-021-29. C.M. is also
2104: supported by Nicolaus Copernicus University Grant No.~408A.
2105: 
2106: \bibliographystyle{mn2e}
2107: \bibliography{ms}
2108: \label{lastpage}
2109: \end{document}
2110: 
2111: 
2112: