astro-ph0007335/ms.tex
1: %\documentclass[manuscript]{aastex}
2: \documentclass[preprint]{aastex}
3: %
4: \newcommand\psr{PSR~B1257+12 }
5: \newcommand\text[1]{\mbox{#1}}                                               
6: \newcommand\mref[1]{(\ref{#1})}
7: \newcommand\figref[1]{Fig.~\ref{#1}}
8: \newcommand{\etal}{{\em et al.}\/\/}
9: \newcommand\pder[2]{\frac{\partial #1}{\partial #2}}
10: \newcommand\der[2]{\frac{d #1}{d #2}}
11: \newcommand\dt{\frac{d \ }{d t}}
12: \newcommand\dder[2]{\frac{d^2 #1}{d {#2}^2}}
13: \newcommand\bmath[1]{\mbox{\protect\boldmath{$#1$}}}
14: \newcommand\vscalar[2]{ \langle \bmath{#1}, \bmath{#2} \rangle}
15: \newcommand\mscalar[2]{ \langle {#1}, {#2} \rangle}
16: \newcommand\vnorm[1]{ \Vert \bmath{#1}\Vert}
17: \newcommand\mnorm[1]{ \Vert  {\bf #1}\Vert}
18: \newcommand\R{{\bf R}}
19: \newcommand\be{{\bf e}}
20: \newcommand\br{{\bf r}}
21: \newcommand\bR{{\bf R}}
22: \newcommand\bb{{\bf b}}
23: \newcommand\bm{{\bf m}}
24: \newcommand\bn{{\bf n}}
25: \newcommand\bk{{\bf k}}
26: \newcommand\bl{{\bf l}}
27: \newcommand\bp{{\bf p}}
28: \newcommand\bq{{\bf q}}
29: \newcommand\bs{{\bf s}}
30: \newcommand\bx{{\bf x}}
31: \newcommand\by{{\bf y}}
32: \newcommand\bz{{\bf z}}
33: \newcommand\bA{{\bf A}}
34: \newcommand\bD{{\bf D}}
35: \newcommand\bE{{\bf E}}
36: \newcommand\bG{{\bf G}}
37: \newcommand\bI{{\bf I}}
38: \newcommand\bJ{{\bf J}}
39: \newcommand\bN{{\bf N}}
40: \newcommand\bL{{\bf L}}
41: \newcommand\bM{{\bf M}}
42: \newcommand\bQ{{\bf Q}}
43: \newcommand\bP{{\bf P}}
44: \newcommand\bS{{\bf S}}
45: \newcommand\bX{{\bf X}}
46: \newcommand\bY{{\bf Y}}
47: \newcommand\bZ{{\bf Z}}
48: \newcommand\bzero{{\bf 0}}
49: %
50: \newcommand\cA{{\cal A}}
51: \newcommand\cB{{\cal B}}
52: \newcommand\cC{{\cal C}}
53: \newcommand\cx{{\rm x}}
54: %
55: \newcommand\eq[1]{(\ref{#1})}
56: \newcommand\rme{{\rm e}}
57: \newcommand\rmi{{\rm i}}
58: 
59: \begin{document}
60: 
61: \title{Improved timing formula for the PSR B1257+12 planetary system}
62: 
63: \author{Maciej Konacki \altaffilmark{1},
64: Andrzej J.~Maciejewski  \altaffilmark{2}}
65: \affil{Toru\'n Centre for Astronomy, 
66:            Nicolaus Copernicus University,\\
67:            87-100 Toru\'n, Gagarina 11, Poland}
68: \author{Alex Wolszczan \altaffilmark{3}}
69: \affil{Department of Astronomy and Astrophysics, Penn State University \\
70: University Park, PA 16802, USA \\
71: Toru\'n Centre for Astronomy, Nicolaus Copernicus University \\
72: ul. Gagarina 11, 87-100 Toru\' n, Poland}
73: \altaffiltext{1}{e-mail: kmc@astri.uni.torun.pl}
74: \altaffiltext{2}{e-mail: maciejka@astri.uni.torun.pl}
75: \altaffiltext{3}{e-mail: alex@astro.psu.edu}
76: 
77: \begin{abstract}
78: We present a new analysis of the dynamics of the planetary system around
79: the pulsar B1257+12. A semi-analytical theory of perturbation between 
80: terrestrial-mass planets B and C is developed and applied to improve
81: multi-orbit timing formula for this object. We use numerical simulations
82: of the pulse arrival times for PSR B1257+12 to demonstrate that our new
83: timing model can serve as a toll to determine the masses of the two planets.
84: \end{abstract}
85: 
86: \keywords{celestial mechanics --- planetary systems --- 
87:           pulsars: individual (PSR B1257+12)} 
88: 
89: \section{Introduction}
90: 
91: The first extra-solar planetary system has been discovered around a 
92: millisecond pulsar B1257+12 \citep{Wolszczan:92::}. The system consists of 
93: three planets named ${\cA}, {\cB}$ and ${\cC}$ with planets ${\cB}$ and 
94: ${\cC}$ having the orbits close to a 3:2 commensurability. This circumstance 
95: allows us to analyze the dynamics of the system beyond the classical Keplerian
96: approximation. Namely, in such configuration, the gravitational interactions 
97: of planets ${\cB}$ and ${\cC}$ give rise to observable time variations of  
98: ${\cB}$ and ${\cC}$ orbital elements. It was thoroughly discussed 
99: by \citet{Rasio:92::} and \citet{Malhotra:93a::} 
100: \citep[see also][]{Malhotra:92::,Malhotra:93b::,Rasio:93::,Peale:93::}. 
101: Subsequently, these studies were used to confirm the existence of the \psr 
102: planetary system through detecting, in the timing observations of the pulsar,
103: the presence of those non-keplerian variations \citep{Wolszczan:94::}
104: 
105: However, the application of non-keplerian dynamics goes further than
106: the confirmation of the discovery.  It can be used to derive some
107: interesting information about the system which is not otherwise
108: accessible.  The aim of this paper is to apply the theory of perturbed
109: planetary motion and derive an improved model for the timing
110: observations of PSR~B1257+12.  Such model, as we show on simulations,
111: should lead to determination of the masses of planets ${\cB}$ and
112: ${\cC}$, as well as the inclinations of their orbits.  To this end, in
113: section \ref{sec:equation} we analyze the equations of motion in the
114: barycentric and Jacobi coordinate systems, which we use in the paper. 
115: In section \ref{sec:timing} we show how these two slightly different representations
116: are related to the commonly used timing model for pulsars with
117: companions. In section \ref{sec:oscul} we demonstrate how to express 
118: such timing formula in terms of the osculating orbital elements. 
119: In section \ref{sec:semi} we show how to obtain the osculating elements of ${\cB}$
120: and ${\cC}$.  In section \ref{sec:timfor} we present an improved timing model
121: describing the motion of this system and finally, in section 
122: \ref{sec:tests}, we perform numerical tests which show how it can be 
123: used in practice.
124:     
125: %
126: \section{Equations of Motion}
127: 
128: \label{sec:equation}
129: 
130: Let us consider a system consisting of a neutron star $P_0$ with the 
131: mass $m_0$, and $N$ planets $P_i$ with masses $m_i$. In an arbitrary 
132: inertial reference frame equations of motion of this system have the form
133: %
134: \begin{equation}
135: \label{eq:nb}
136: m_i \ddot{\bR}_i = -G\sum_{ {\stackrel{\scriptstyle j=0}{j\neq i}}}^N
137: \frac{ m_i m_j}{R_{ij}^3}\left(\bR_i-\bR_j\right),
138: \end{equation}
139: %
140: where
141: \[
142: R_{ij}=\Vert\bR_i-\bR_j\Vert, \qquad i, j = 0, \ldots, N,
143: \]
144: %
145: and $G$ is the gravitational constant. The Hamiltonian function for system 
146: \mref{eq:nb} has the form
147: %
148: \begin{equation}
149: \label{eq:h}
150: H = \frac{1}{2}\sum_{i=0}^N\frac{1}{m_i}\bP_i^2 - \sum_{0\leq i<j\leq N}
151: \frac{G m_im_j}{R_{ij}},
152: \end{equation}
153: %
154: and equations \mref{eq:nb} can be written as Hamilton's equations
155: %
156: \begin{equation}
157: \label{eq:he}
158: \dt{\bR_i} = \pder{H}{\bP_i}, \qquad \dt{\bP_i} = -\pder{H}{\bR_i}, 
159: \qquad i = 0, \ldots, N.
160: \end{equation}
161: %
162: Analytical perturbation theory for a planetary system is usually formulated 
163: in the so-called Jacobi coordinates $\br_i$, $i=0, \ldots, N$ which are 
164: defined in the following way
165: %
166: \begin{equation}
167: \label{eq:jac}
168: \br_k = \bR_k - \frac{1}{\mu_{k-1}}\sum_{i=0}^{k-1} m_i\bR_i,\quad\text{for}
169: \quad k=1,\ldots, N,
170: \end{equation}
171: %
172: and 
173: %
174: \begin{equation}
175: \label{eq:jac0}
176: \br_0 =  \frac{1}{\mu_{N}}\sum_{i=0}^{N} m_i\bR_i,\qquad \mu_k=
177: \sum_{i=0}^km_i.
178: \end{equation}
179: %
180: In other words, $\br_i$ is the radius vector from the
181: center of mass of bodies $P_0, \ldots, P_{i-1}$ to body $P_i$, and $\br_0$ 
182: is the center of mass of the system. The above formulae define a one-to-one 
183: relationship between Cartesian inertial coordinates and Jacobi coordinates. 
184: The inverse relationship has the form
185: %
186: \begin{equation}
187: \label{eq:jaci}
188: \bR_k = \br_0 + \frac{\mu_{k-1}}{\mu_k}\br_k - \sum_{i=k+1}^N\frac{m_i}{\mu_i}
189: \br_i, \qquad k=0,\ldots, N,
190: \end{equation}
191: % 
192: where we assume $\mu_{-1}=0$.
193: %  
194: In terms of canonical Jacobi coordinates, Hamiltonian \mref{eq:h} reads
195: %
196: \begin{equation}
197: \label{eq:hj}
198: \begin{array}{l}
199: \displaystyle
200: H = \frac{1}{\mu_N}\bp_0^2 + \sum_{i=1}^N\frac{\mu_i}{\mu_{i-1}m_i}\bp_i^2
201:   - G\sum_{i=1}^{N}\frac{m_0m_i}{r_i} \; + \\[0.5cm]
202: \displaystyle
203: \qquad + \;\tilde H_1\left(\br_1,\ldots, \br_N\right),
204: \end{array}
205: \end{equation}
206: %
207: where $\tilde H_1=\tilde H_1(\br_1,\ldots, \br_N)$ is its perturbative part. 
208: If we assume that masses of planets are small and all 
209: are of the same order $\epsilon$, then $\tilde H_1$ is of order 
210: $\epsilon^2$, i.e., $\tilde H_1 = H_1 + {\cal O}(\epsilon^3)$, where
211: %
212: \begin{equation}
213: \label{eq:h1}
214: H_1 = -G\!\!\!\sum_{1\leq i< j\leq N} \!\!\!m_i m_{j}\left[ \frac{1}{r_{ij}} 
215: - \frac{\br_i\cdot\br_j}{r_i}\right].
216: \end{equation}
217: % 
218: From the form of  Hamiltonian \mref{eq:hj} it follows that $\bp_0$ is a first 
219: integral and that the equations for variables 
220: $\{\br_1,\ldots, \br_N, \bp_1, \ldots, \bp_N\}$ do not depend on its 
221: particular value. Thus, we can assume that $\bp_0=\bzero$ implying 
222: that $\br_0$ is constant and can be set to $\br_0=\bzero$. 
223: This is equivalent to the assumption that our inertial frame is a certain 
224: barycentric reference frame. The dynamics of the system is governed by 
225: Hamilton's equations with the Hamiltonian of the form  
226: %
227: \begin{equation}
228: \label{eq:hper}
229: H = H_0 + H_1 + \cdots,
230: \end{equation}
231: %
232: where 
233: %
234: \begin{equation}
235: \label{eq:h0}
236: H_0 = \sum_{i=1}^N\frac{\mu_i}{\mu_{i-1}m_i}\bp_i^2  - G\sum_{i=1}^{N}
237: \frac{m_0m_i}{r_i},
238: \end{equation}
239: %
240: is the unperturbed part of the Hamiltonian describing a system of $N$ 
241: independent planets. It follows from the form of Eq. \mref{eq:h0} that 
242: each planet moves in a Keplerian orbit in the same way as a body with the mass 
243: $m_i\mu_{i-1}/\mu_i$ around a fixed gravitational center $m_0 m_i$. 
244: Each of these Keplerian motions can be parameterized by Keplerian elements 
245: $\{ T_p, a,e,i,\omega,\Omega\}$---the time of pericenter, semi-major axis, 
246: eccentricity, inclination, argument of pericenter  and the longitude of
247: ascending node, respectively.
248: 
249: \section{Timing model and coordinate systems}
250: 
251: \label{sec:timing}
252: 
253: Timing observations of pulsars represent measurements of the times of arrival
254: of pulsars pulses (TOAs). An extraordinary precision of timing measurements 
255: allows a detection of very low-level effects in timing residuals 
256: \citep[see for review][]{Lyne:98::}. In the case of a binary pulsar the 
257: observed TOAs exhibit periodic variations resulting from the motion of the 
258: pulsar around the center of mass. Such variations are modeled with the 
259: formula
260: %
261: \begin{equation}
262: \Delta t = x\left[\left(\cos E - e\right)\sin\omega + \sqrt{1 - e^2}\sin
263: E\cos\omega\right],
264: \end{equation}
265: where
266: %
267: \[
268: x = a\sin i/c, \quad E - e \sin E = n\left(t- T_p\right), \quad
269: n = \frac{2\pi}{P}, 
270: \]
271: %
272: and $\Delta t$ is the additional delay/advance in TOAs, $E$ is eccentric
273: anomaly, ${a,e,\omega,\sin i,P,T_p}$ are Keplerian elements of the orbit
274: and $c$ is the speed of light. When a pulsar has $N$ planets the
275: TOA variations become
276: %
277: \begin{equation}
278: \begin{array}{l}
279: \label{eq:kep::}
280: \displaystyle
281: \Delta t = \sum_{j=1}^{N}x_j\Bigl[\left(\cos E_j - e_j\right)\sin\omega_j 
282: \; + \\[0.5cm]
283: \displaystyle
284: \hspace{2cm} + \; \left.\sqrt{1 - e_j^2}\left.\sin E_j\cos\omega_j
285: \right)\right].
286: \end{array}
287: \end{equation}
288: %
289: Note that in practice the parameters of the model which we  
290: determine, by means of the least-squares fit to the data, are 
291: ${x_j,e_j,\omega_j,P_j,T_{pj}}$, i.e., the projection of  semi-major axis
292: of the {\it pulsar's} orbit, 
293: eccentricity, argument of pericenter, orbital period and time of pericenter. 
294: In order to precisely understand and interpret these parameters we describe 
295: the pulsar's motion in the barycentric reference frame with
296: the $z$-axis of the system directed toward the barycenter of the solar system 
297: and $xy$ plane of the reference frame in the plane of the sky. This way, 
298: we have
299: %
300: \begin{equation}
301: \label{eq:dt}
302:   \Delta t = -\frac{1}{c}  \bR_0\cdot \widehat\bZ ,  
303:   \quad \bR_0 = - \frac{1}{m_0} \sum_{i=1}^N m_i\bR_i,
304: \end{equation}
305: %
306: where $\widehat\bZ$ is the unit vector along the $z$-th axis of the pulsar
307: system barycentric frame and $\bR_j$ are barycentric positions of planets. 
308: Assuming that
309: %
310: \begin{equation}
311: \frac{m_j}{m_0}a_j\sin i_j = x_j c, 
312: \quad \frac{Gm_0}{(1 + m_j/m_0)^2} = n_j^2 a_j^3,
313: \end{equation}
314: %
315: with {\it planets'} orbital parameters $a_j,n_j,e_j,\omega_j,T_{pj}$, 
316: we can most naturally interpret the motion of the pulsar as a superposition 
317: of the elliptic motions of its planets around the barycenter of the system. 
318: However, as it was mentioned in the previous section, the analytical 
319: perturbation theory is usually formulated in Jacobi coordinates in which 
320: the TOA variations become
321: %
322: \begin{equation}
323: \Delta t = -\frac{1}{c}  \bR_0\cdot \widehat\bZ, \quad  
324: \bR_0 = -  \sum_{j=1}^N \kappa_j\br_j, \qquad
325: \kappa_j = \frac{m_j}{\mu_j}, \quad
326: \end{equation}
327: %
328: where $\br_j$ are positions of planets. Furthermore, we have the following 
329: relations
330: %
331: \begin{equation}
332: \kappa_ja_j\sin i_j = x_j c, \quad 
333: \frac{Gm_0\mu_j}{\mu_{j-1}} = \frac{Gm_0}{1 - \kappa_j}  
334: = n_j^2 a_j^3. 
335: \end{equation}
336: %
337: Thus from the timing formula for TOA variations of a pulsar with planets
338: it is possible to obtain two somewhat different descriptions of the pulsar
339: motion. Although, they both represent a sum of a certain number of elliptic 
340: motions, the interpretation of some of their parameters is slightly 
341: different. Throughout the rest of this paper, we will use Jacobi coordinates 
342: as they are more convenient in the formulation of the theory of perturbed 
343: motion.
344: %
345: 
346: \section{Osculating orbital elements}
347: %
348: \label{sec:oscul}
349: The non-keplerian motion of the \psr system can be described by 
350: means of the osculating ellipses (i.e. by means of ellipses which
351: parameters change with time). The time evolution of orbital elements
352: can be determined by solving the classical Lagrange's perturbation
353: equations \citep[see, for example][]{Brouwer:61::} with the perturbation
354: Hamiltonian given by equation \mref{eq:h1}. Such approach leads
355: to the solution in the form 
356: %
357: \begin{equation}
358: \label{eq:dx}
359: \cx = \cx^0 + \Delta\cx\left(t-t_0\right),
360: \end{equation}
361: %
362: where $\cx$ stands for a specific orbital element, $\cx^0$ its initial 
363: value and $\Delta\cx$ for its time dependent part of small magnitude 
364: $\Delta\cx(t-t_0)/\cx^0 \ll 1$. In the case of the \psr planetary
365: systems the most significant part of the perturbations comes from
366:  planets $\cB$ and $\cC$. Therefore, we 
367: can assume that the orbital elements of  planet $\cA$ are approximately 
368: constant while the elements of planets $\cB$ and $\cC$ change with time.
369: Thus using the formulae for $\Delta t$ from the previous section 
370: (Eq. \mref{eq:kep::}) and assuming the time evolution of orbital elements in 
371: the form \mref{eq:dx}, the additional TOA variations 
372: $\delta t_j, j = \{\cB,\cC\}$ due to the interactions between 
373: planets $\cB$ and $\cC$ can be expressed as follows
374: %
375: \begin{equation} 
376: \label{eq:small}
377: \begin{array}{l}
378: \displaystyle
379: \frac{c\delta t_j}{\kappa_j\sin i^0_j} = 
380: -\Delta h_j\left(\frac{3}{2}a_j^0 + \frac{1}{2}a_j^0
381: \cos\left(2\lambda^0_j\right)\right) \; +
382: \\[0.3cm] 
383: \displaystyle
384: \hspace{2cm} + \; \frac{1}{2}\Delta k_j a_j^0\sin\left(2\lambda^0_j\right) 
385: \; + \\[0.3cm]
386: \displaystyle
387: \hspace{2cm} + \; \Delta a_j\sin\lambda_j^0 + 
388: \Delta\lambda_ja_j^0\cos\lambda^0_j,
389: \end{array} 
390: \end{equation}
391: %
392: where
393: %
394: \begin{equation}
395: \begin{array}{l}
396: \displaystyle
397: h_j = e_j\sin\omega_j, \quad k_j = e_j\cos\omega_j, \\[0.3cm]
398: \displaystyle
399: \lambda_j = n_j(t - T_{pj}) + \omega_j,
400: \end{array}
401: \end{equation} 
402: %
403: and equation \mref{eq:small} is given to first order in 
404: $\Delta a_j, \Delta\lambda_j, \Delta h_j, \Delta k_j$ and
405: the lowest order in $e_j$ (or $h_j$ and $k_j$, since
406: the eccentricities of planets $\cB$ and $\cC$ are very small).
407: The above equations can be obtained from the following 
408: well-known expansions
409: %
410: \begin{equation}
411: \begin{array}{c}
412: \displaystyle
413: \cos E = -\frac{e}{2} + 2\sum_{k \in {\cal Z}_0}\frac{1}{k}J_{k-1}(ke)\cos(k M),
414: \\[0.3cm]
415: \displaystyle
416: \sin E = \frac{2}{e}\sum_{k \in {\cal Z}_0}\frac{1}{k}J_{k-1}(ke)\sin(k M),
417: \quad M = \lambda - \omega
418: \end{array}
419: \end{equation}
420: %
421: where $J_k(x)$ is a Bessel function of the first kind of order $n$ and
422: argument $x$, ${\cal Z}_0$ denotes the set of all positive and negative integers  
423: excluding zero; and an approximate relation for Bessel functions of 
424: small arguments
425: %
426: \begin{equation}
427: J_{k}(x) \approx \frac{x^k}{2^k k!}, \quad x << 1
428: \end{equation}
429: %
430: 
431: Now, our next step is to find the explicit form of the functions
432: %
433: \begin{equation}
434: \begin{array}{l}
435: \displaystyle
436: a_j(t) = a_j^0 + \Delta a_j(t), \quad \lambda_j(t) = \lambda_j^0 + 
437: \Delta \lambda_j(t), \\[0.3cm]
438: \displaystyle
439: h_j(t) = h_j^0 + \Delta h_j(t), \quad k_j(t) = k_j^0 + \Delta k_j(t).
440: \end{array}
441: \end{equation}
442: %
443: 
444: 
445: \section{Semi-analytical perturbation theory}
446: 
447: \label{sec:semi}
448: Mutual gravitational interactions between planets cause periodic and 
449: secular changes of their orbital elements. In the case of  \psr
450:  periodic variations can be related to conjunctions (`close encounters') 
451: of  planets $\cB$ and $\cC$ with the frequency $n_{ce} = n_\cB - n_\cC$
452: and to the 3:2 near-commensurability with the frequency 
453: $n_{r} = 3n_{\cC}-2n_{\cB}$. The dynamics of this system was
454: studied by \citet{Rasio:92::} and \citet{Malhotra:93a::} \citep[see 
455: also][]{Malhotra:93b::,Rasio:93::,Peale:93::}. \citet{Malhotra:93a::} 
456: solved the Lagrange's perturbation equations
457: for the orbital elements assuming non-resonant system with coplanar orbits. 
458: This first order perturbation theory is valid for $(\sin i)^{-1} \gtrsim 10$ 
459: (orbital inclinations larger than about 6 degrees) which corresponds to mass 
460: ratios $m_j/m_0 \gtrsim 6\times10^{-5}$ (and non-resonant case). When  
461: planets $\cB$ and $\cC$ are locked in the exact resonance it is necessary 
462: to develop a different theory of motion \citep{Malhotra:92::,Rasio:93::}.
463: 
464: It turns out that the first-order perturbation theory for this system
465: developed by \citet{Malhotra:93a::} is not accurate enough for 
466: the purpose
467: of determination of the planets' masses. Therefore, in this paper we present
468: a new approach to this problem which precisely addresses the issue.
469: Namely, first we assume that the orbits of ${\cB}$ and ${\cC}$ are not
470: coplanar however their relative inclination $I$ is small. The geometry
471: of the system can be represented as in Fig. 1 and the perturbative part of
472: Hamiltonian expanded to the first order in $I$ and the product of the
473: mass of planets ${\cB}$ and ${\cC}$, $m_1m_2$, has the following form
474: %
475: \begin{equation}   
476: H_1 = -\frac{Gm_1m_2}{r_2}\left[\left(1 - 2\frac{r_1}{r_2}
477: \cos\psi 
478: + \left(\frac{r_1}{r_2}\right)^2\right)^{-1/2} -\frac{r_1}{r_2}\cos\psi\right],
479: \end{equation}
480: %
481: where
482: %
483: \[
484: \begin{array}{l}
485: \displaystyle
486: \psi = f_1 + \omega_1 - f_2 - \omega_2 - \tau, \quad 
487: \tau = \tau_1 - \tau_2 ,\\[0.5cm]
488: \displaystyle
489: r_j = \frac{a_j\left(1 - e_j^2\right)}
490: {1 + e_j\cos f_j}, \quad j={1,2},
491: \end{array}
492: \]
493: %
494: and $f_j$ is the true anomaly. Note that in the above expansion,
495: $H_1$ does not depend on $I$. The first term in which $I$ appears is 
496: of the second order; precisely it is $\sin^2(I/2)$. Thus as long as
497: $\sin^2(I/2)$ is negligible, the above expansion is valid.
498: 
499: Next, from the classical form of Lagrange's perturbation equations 
500: \citep[see, for example][]{Brouwer:61::} we can obtain the following set 
501: of the first order differential equations for the elements $a_j, e_j, 
502: \omega_j, \lambda_j$
503: %
504: \begin{equation}
505: \label{osc}
506: \begin{array}{c}
507: \displaystyle
508: \dot a_j = \frac{-2}{\mu_jn_ja_j}
509: \frac{\partial H_1}{\partial\sigma_j}, \\[0.7cm]
510: \end{array}
511: \end{equation}
512: \[
513: \begin{array}{c}
514: \displaystyle
515: \dot e_j = \frac{1}{\mu_jn_ja_j^2e_j}\left[
516: -\left(1 - e_j^2\right)\frac{\partial H_1}{\partial\sigma_j}
517: + \sqrt{1 - e_j^2}\frac{\partial H_1}{\partial\omega_j}\right], \\[0.7cm]
518: \displaystyle
519: \dot\omega_j = 
520:  \frac{-\sqrt{1 - e_j^2}}{\mu_jn_ja_j^2e_j}
521: \frac{\partial H_1}{\partial e_j}, \\[0.7cm]
522: \displaystyle
523: \dot\lambda_j = n_j + \frac{2}{\mu_jn_ja_j}\frac{\partial H_1}{\partial a_j}
524: - \frac{e_j\sqrt{1 - e_j^2}}{\mu_jn_ja_j^2\left(1 + \sqrt{1 - e_j^2}\right)}
525: \frac{\partial H_1}{\partial e_j},
526: \end{array}
527: \]
528: %
529: where $n_j$ is the mean motion and $\sigma_j$ is related to time of
530: pericenter through $\sigma_j = -n_j T_{pj}$ so as the Kepler equation
531: reads 
532: %
533: \begin{equation}
534: E_j - e_j\sin E_j = n_jt + \sigma_j = \lambda_j - \omega_j,
535: \end{equation}
536: %
537: and the true anomaly $f_j$ is related to the eccentric anomaly $E_j$
538: through 
539: %
540: \begin{equation}
541: \tan\frac{f_j}{2} = \sqrt{\frac{1 + e_j}{1 - e_j}}\tan\frac{E_j}{2}.
542: \end{equation}
543: %
544: And it should be noted that the elements $\Omega_j, i_j$ remain approximately
545: constant in the case of small relative inclinations. 
546: 
547: In principle, such set of equations could be solved for $a_j, e_j, \omega_j,
548: \lambda_j$. In practice however, it is extremely complicated. On the other 
549: hand, in fact we do not need analytical formulae such as those presented in 
550: \citet{Malhotra:93a::}. Thus the most suitable approach is to solve 
551: this problem numerically. In order to do so we just need the explicit form
552: of the right-hand-side functions of  equations \mref{osc} which can be
553: easily obtained using the perturbing Hamiltonian $H_1$. Subsequently, 
554: the equations can be solved numerically for $a_j, e_j, \omega_j, \lambda_j$. 
555: As we show in section 7, this approach gives very accurate results, much 
556: more accurate than any reasonable analytical treatment. From the form of 
557: equations \mref{osc} it follows that the change of  orbital parameters 
558: $\Delta\cx$ (strictly speaking here $\Delta\cx$ stands for 
559: $\Delta\lambda,\Delta h,\Delta k$ and $\Delta a/a_0$) is proportional to 
560: $m_1/m_0$ and $m_2/m_0$. Precisely, corrections $\Delta\cx_1$ are proportional
561: to $m_2/m_0$ and $\Delta\cx_2$ are proportional to $m_1/m_0$. Let us finally 
562: note that in this model there is one additional parameter $\tau$ which in 
563: general is not known a priori. However, it can be determined through a 
564: least-squares analysis of the data. 
565: 
566: \section{Timing formula}
567: \label{sec:timfor}
568: 
569: We have now all the components necessary to derive a useful timing formula
570: which will describe the motion of the \psr system. First of all, let us note
571: that due to the small mass of planet ${\cal A}$ we will have:
572: %
573: \begin{equation}
574: \begin{array}{c}
575: \displaystyle
576: \kappa_{\cA} = \kappa_1 = \frac{m_{\cA}}{m_0}, \\[0.4cm]
577: \displaystyle
578: \kappa_{\cB} = \kappa_2 = \frac{m_{\cB}}{m_0 + m_{\cA} + m_{\cB}}
579: \approx \frac{m_{\cB}}{m_0 + m_{\cB}}, \\[0.4cm] 
580: \end{array}
581: \end{equation}
582: \[
583: \kappa_{\cC} = \kappa_3 = \frac{m_{\cC}}{m_0 + m_{\cA} + m_{\cB} + m_{\cC}}
584: \approx \frac{m_{\cC}}{m_0 + m_{\cB} + m_{\cC}}. 
585: \]
586: %
587: 
588: The problem of finding masses and inclinations of planets ${\cB}$ and
589: ${\cC}$ can be now formulated as a least-squares problem in which we fit
590: to the data the following function
591: %
592: \begin{equation}
593: \Delta t = \Delta t_{K}(\Psi) + 
594: \delta t_{\cB}(\Psi,\gamma_{\cB},\gamma_{\cC}) + 
595: \delta t_{\cC}(\Psi,\gamma_{\cB},\gamma_{\cC}),
596: \end{equation}
597: where 
598: \begin{equation}
599: \begin{array}{l}
600: \displaystyle
601: \Psi = \left\{x^0_{\cB},n^0_{\cB},e^0_{\cB},\omega^0_{\cB},
602: T^0_{p\cB},x^0_{\cC},n^0_{\cC},e^0_{\cC},\omega^0_{\cC},
603: T^0_{p\cC}\right\}, \\[0.4cm]
604: \displaystyle
605: \gamma_{\cB} = \frac{m_{\cB}}{m_0}, \quad \gamma_{\cC} =
606: \frac{m_{\cC}}{m_0},
607: \end{array}
608: \end{equation}
609: %
610: $\Delta t_{K}(\Psi)$ describes the Keplerian part of the motion given
611: by  equation \mref{eq:kep::} and
612: %
613: \begin{equation}                                      
614: \begin{array}{l}
615: \displaystyle
616: \delta t_{\cB} = x_{\cB}^0\Bigg[-\Delta h_{\cB}\left(\frac{3}{2} 
617: + \frac{1}{2}\cos\left(2\lambda^0_{\cB}\right)\right) \; +
618: \\[0.4cm] 
619: \displaystyle
620: \hspace{2cm} + \; \frac{1}{2}\Delta k_{\cB}\sin
621: \left(2\lambda^0_{\cB}\right) 
622: \; + \\[0.4cm]
623: \displaystyle
624: \hspace{2cm} + \; \frac{\Delta a_{\cB}}{a_{\cB}^0}\sin\lambda_{\cB}^0 + 
625: \Delta\lambda_{\cB}\cos\lambda^0_{\cB}\Bigg],\\[0.5cm]
626: %
627: \displaystyle
628: \delta t_{\cC} = x_{\cC}^0\Bigg[-\Delta h_{\cC}\left(\frac{3}{2} 
629: + \frac{1}{2}\cos\left(2\lambda^0_{\cC}\right)\right) \; +
630: \\[0.4cm] 
631: \displaystyle
632: \hspace{2cm} + \; \frac{1}{2}\Delta k_{\cC}\sin
633: \left(2\lambda^0_{\cC}\right) 
634: \; + \\[0.4cm]
635: \displaystyle
636: \hspace{2cm} + \; \frac{\Delta a_{\cC}}{a^0_{\cC}}
637: \sin\lambda_{\cC}^0 + 
638: \Delta\lambda_{\cC}\cos\lambda^0_{\cC}\Bigg],
639: \end{array}
640: \end{equation}
641: %
642: describe its non-keplerian part. In this model the initial values of 
643: osculating orbital
644: elements replace the Keplerian elements as parameters of the model
645: and we additionally have the parameters, $\gamma_{\cB}$ and 
646: $\gamma_{\cC}$, related to the masses of ${\cB}$ and ${\cC}$. We also
647: need the derivatives of $\Delta t$ with respect to the parameters 
648: $\{\Psi,\gamma_{\cB},\gamma_{\cC}\}$. With sufficient accuracy the
649: derivatives with respect to $\Psi$ can be computed from the Keplerian part
650: $\Delta t_K$ of the model only and the derivatives with respect to $\gamma_{\cB}$ and 
651: $\gamma_{\cC}$ can be easily obtained as $\delta t_{\cB}$ and
652: $\delta t_{\cC}$ are proportional to them
653: %
654: \begin{equation}
655: \frac{\partial\Delta t}{\partial{\gamma_{\cB}}} = 
656: \frac{\delta t_{\cC}}{\gamma_{\cB}}, \quad
657: \frac{\partial\Delta t}{\partial{\gamma_{\cC}}} = 
658: \frac{\delta t_{\cB}}{\gamma_{\cC}}.
659: \end{equation}
660: %
661:  
662: Now, determination of masses of ${\cB}$ and ${\cC}$ and inclinations 
663: $i_{\cB}$ and $i_{\cC}$ can be carried out in the following way. First, we 
664: assume that the mass of the pulsar $m_0$ is canonical (i.e. $1.4$M$_{\odot}$) 
665: and derive $m_{\cB}$, $m_{\cC}$ directly from $\gamma_{\cB}$ and $\gamma_{\cC}$.
666: Subsequently inclinations can be computed from the following equations
667: %
668: \begin{equation}
669: c x^0_j = \kappa_j a^0_j\sin i^0_j,\quad
670: \frac{Gm_0}{1 - \kappa_j} = {n^0_j}^2{a^0_j}^3,
671: \end{equation}
672: %
673: where $j = \{\cB,\cC\}$.
674: 
675: \section{Numerical tests and conclusions}
676: 
677: \label{sec:tests}
678: 
679: We start the tests with the comparison of our approach for computing
680: osculating elements with that of \citet{Malhotra:93a::}. It turns
681: out that the most significant component of the non-keplerian motion comes
682: from the changes of mean longitude $\lambda_j$. Therefore in Fig. 2
683: we present $\Delta\lambda_j$ for the case of coplanar, edge-on orbits.
684: As we can see, the approach developed in this paper is in fact as good
685: as the integration of full equations of motion. For other elements
686: we obtain very similar results. It should be mentioned that because 
687: the model of \citet{Malhotra:93a::} is not accurate  enough in 
688: predicting the secular change of $\Delta\lambda_j$ and because
689: $\Delta\lambda_j$ are proportional to $\gamma_{\cB}$ and $\gamma_{\cC}$, using 
690: such model would result in a significant error in determination of the 
691: planets' masses. Next, we compared the TOA residuals due to the non-keplerian
692: part of the motion calculated from our model with those obtained by means
693: of  numerical integration of full equations of motion. In Fig. 3,
694: as one can see, for small relative inclinations (Fig. 3 (a) and (b)) our 
695: model is very accurate. For larger $I$ (Fig. 3 (c) and (d)) one can 
696: see small differences but this is consistent with the assumptions made 
697: to obtain the model.
698:   
699: Finally, we performed the tests which show how the derived timing model
700: can be used to obtain the masses of planets. To this end we used the 
701: TEMPO software package (see the Internet location 
702: {\tt http://pulsar.princeton.edu/tempo}) modified to include our model
703: of the non-keplerian motion of  \psr. We simulated two different
704: sets of artificial TOAs assuming the parameters of  \psr
705: \citep{Wolszczan:94::}. The first set of TOAs sampled every day 
706: covered a period of 10 years. We also added Gaussian noise with
707: $\sigma = 0.1 \mu s$. The second set was prepared to resemble the real
708: timing observations of PSR B1257+12. Under such conditions we simulated
709: TOAs for the cases described in Fig. 3. Subsequently, we applied
710: the modified TEMPO to obtain the masses of planets $\cB$ and $\cC$. The
711: results are presented in Table 1.
712: 
713: The tests performed on  Set 1 were used to establish a limit on 
714: applicability of our model. As one can see, for the relative inclinations
715: $I$ of about 10 degrees the model becomes inaccurate and the relative error
716: of determination of masses is at the level of 20$\%$. Because the relative
717: inclination weakens the interactions of planets, using the model which
718: assumes a small $I$, results, in such situations, in the determined masses 
719: that are smaller than the real ones. From the tests performed on  Set 2 
720: we also learn that this error is bigger than uncertainty originating in
721: TOA measurement errors. On the other hand, when $I$ is relatively small 
722: we obtain a very accurate determination of  masses thus proving that
723: our model can be successfully applied as long as the assumptions made
724: to derive it are satisfied.
725: 
726: To sum up, our model can be applied in all
727: the cases when the `observational' uncertainties of the planets' masses
728: are bigger than the uncertainties resulting from the assumptions made to
729: derive the timing model. This, in general, depends on  masses of planets
730: and the relative inclination of the orbits $I$ as well as the
731: characteristic of the observations  but one can estimate that
732: in the case of the \psr data $I$ should be smaller than 10 degrees.
733: We should also mention that for non-coplanar orbits the angle
734: $\tau$ will be in general different from zero therefore we have to find
735: it in order to get the proper values of the planets' masses. It can 
736: be done by computing the least-squares value of $\chi^2$ for a range of 
737: $\tau$ and then choosing the $\tau$ that corresponds to the minimum
738: of $\chi^2$. And eventually, one has to remember that the \psr planetary
739: system could be in such configuration that a more significant alteration
740: must be done to our model. First of all, the relative inclination
741: of the orbits could be large. In such  case one would have to use
742: the full form of the perturbative Hamiltonian $H_1$ which would lead
743: to a much more complicated model with four additional parameters $\Omega_{\cB},
744: \Omega_{\cC},i_{\cB},i_{\cC}$ (instead of one $\tau$). Secondly, the 
745: presence of a massive distant planet, if significant for the theory of 
746: motion, would have to be taken into account.
747: 
748: \begin{thebibliography}{}
749: 
750: \bibitem[Brouwer \& Clemence (1961)]{Brouwer:61::} Brouwer, D.  \& 
751: Clemence, G. M. 1961, Methods of Celestial Mechanics,
752: (New York: Academic Press)
753: 
754: \bibitem[Lyne \& Graham-Smith (1998)]{Lyne:98::} Lyne, A. G. \& 
755: Graham-Smith, F. 1998, Pulsar Astronomy, 
756: (New York: Cambridge University Press)
757: 
758: \bibitem[Malhotra \etal (1992)]{Malhotra:92::} Malhotra, 
759: R., Black, D., Eck, A. \& Jackson, A. 1992, \nat, 356, 583 
760: 
761: \bibitem[Malhotra (1993a)]{Malhotra:93a::} Malhotra, R. 1993a, \apj, 407, 
762: 266 
763: 
764: \bibitem[Malhotra (1993b)]{Malhotra:93b::} Malhotra, R. 1993b, ASP Conf. 
765: Ser. 36: Planets Around Pulsars, 89, (San Francisco: ASP) 
766: 
767: \bibitem[Peale (1993)]{Peale:93::} Peale, S. J. 1993, \aj, 105, 1562 
768: 
769: \bibitem[Rasio \etal (1992)]{Rasio:92::} 
770: Rasio, F. A., Nicholson, P. D., Shapiro, S. L. \& Teukolsky, S. A. 1992, 
771: \nat, 355, 325 
772: 
773: \bibitem[Rasio \etal (1993)]{Rasio:93::} 
774: Rasio, F. A., Nicholson, P. D., Shapiro, S. L. \& Teukolsky, S. A. 1993, 
775: ASP Conf. Ser. 36: Planets Around Pulsars, 107, (San Francisco: ASP) 
776: 
777: \bibitem[Wolszczan \& Frail (1992)]{Wolszczan:92::} Wolszczan, A. \&
778: Frail, D. A. 1992, \nat, 355, 145
779: 
780: \bibitem[Wolszczan (1994)]{Wolszczan:94::}
781: Wolszczan, A. 1994, Science, 264, 538
782: 
783: \end{thebibliography}
784: 
785: \clearpage
786: 
787: %
788: % FIGURE CAPTIONS
789: %
790: 
791: \figcaption[fig1.ps]
792: {Geometry of the system. The angles $\omega_1$, $\omega_2$ 
793: are the arguments of pericenter, $\Omega_1$, $\Omega_2$ longitudes 
794: of ascending node and $i_1$, $i_2$ inclinations of the orbit of the 
795: planets ${\cB}$ and ${\cC}$; $\tau_1$ and $\tau_2$ are the angles
796: $n_1On_{12}$ and $n_2On_{12}$ respectively. The angles $I,\tau_1,\tau_2$
797: can be found by solving the spherical triangle $n_1n_{12}n_2$.}
798: 
799: \figcaption[fig2.ps]
800: {Time evolution of the mean longitude 
801: $\Delta\lambda(t) = \lambda(t) - \lambda^0(t)$ for 
802: the planets $\cB$ and $\cC$ in the case of coplanar, edge-on orbits. 
803: The solid line indicates the solution
804: by means of the numerical integration of full equations of motion,
805: the open circles indicate the solution obtained with the approach
806: presented in this paper and the dash-dotted line with the model by
807: \cite{Malhotra:93a::}.}
808: 
809: \figcaption[fig3.ps]
810: {TOA residuals due to the non-keplerian part of motion of the 
811: PSR B1257+12 in four different configurations. The solution
812: obtained by means of the numerical integration of full equations
813: of motion is indicated with the dots and the one obtained using
814: our model with the solid line.}
815: 
816: \clearpage
817: 
818: %
819: % TABLES
820: %
821: 
822: %
823: % TABLE 1
824: %
825: 
826: \begin{deluxetable}{lrrrrrrrrrr}
827: \tablecolumns{11}
828: \tablewidth{0pc}
829: %
830: %\tablewidth{460pt}
831: %\scriptsize
832: \tablecaption{ Assumed and derived parameters of simulations
833: \tablenotemark{a}}
834: %
835: \tablehead{
836: \colhead{} &
837: \multicolumn{4}{c}{Assumed} &
838: \colhead{} &
839: \multicolumn{2}{c}{Set 1} &
840: \colhead{} &
841: \multicolumn{2}{c}{Set 2} \\
842: \cline{2-5} \cline{7-8} \cline{10-11}\\
843: \colhead{} &
844: \colhead{$\tau$} &
845: \colhead{$I$} &
846: \colhead{$m_{\cB}$} &
847: \colhead{$m_{\cC}$} &
848: \colhead{} &
849: \colhead{$m_{\cB}$} &
850: \colhead{$m_{\cC}$} &
851: \colhead{} &
852: \colhead{$m_{\cB}$} &
853: \colhead{$m_{\cC}$} \\
854: \colhead{} & \colhead{} & \colhead{} & 
855: \colhead{$[M_{\earth}]$} & \colhead{$[M_{\earth}]$} &
856: \colhead{} &
857: \colhead{$[M_{\earth}]$} & \colhead{$[M_{\earth}]$} &
858: \colhead{} &
859: \colhead{$[M_{\earth}]$} & \colhead{$[M_{\earth}]$} }
860: \startdata
861: (a)\dotfill& $0.\!\!^\circ0$ & $0.\!\!^\circ0$ & 3.41 & 2.83 & 
862: & 3.42(1) & 2.82(1) & &
863: 3.53(51) & 2.63(51) \\
864: (b)\dotfill& $0.\!\!^\circ0$ & $2.\!\!^\circ0$ &4.99 & 4.32 &   
865: & 4.98(1) & 4.26(1) & & 
866: 5.07(51)& 4.06(48) \\
867: (c)\dotfill& $0.\!\!^\circ0$ & $10.\!\!^\circ0$& 4.82 & 4.94 &  
868: & 4.06(3) & 4.11(3) &  &
869: 4.06(51)&3.79(48) \\
870: (d)\dotfill& $9.\!\!^\circ7$ & $10.\!\!^\circ3$ &9.96 & 16.33 &  
871: & 8.42(12) & 13.39(12)& &
872: 8.25(45)&13.16(42) \\
873: \enddata
874: \tablenotetext{a}{Figures in parentheses are $3\sigma$
875: uncertainties in the last digits quoted.}
876: \end{deluxetable}
877: %
878: 
879: \clearpage
880: 
881: %
882: % FIGURES
883: %
884: 
885: %
886: % FIGURE 1
887: %
888: 
889: \begin{figure}
890: \figurenum{1}
891: \epsscale{0.75}
892: \plotone{fig1.ps}
893: \caption{}
894: \end{figure}
895: %
896: 
897: 
898: %
899: % FIGURE 2
900: %
901: 
902: \begin{figure}
903: \figurenum{2}
904: \epsscale{0.9}
905: \plotone{fig2.ps}
906: \caption{}
907: \end{figure}
908: %
909: 
910: 
911: %
912: % FIGURE 3
913: %
914: 
915: \begin{figure}
916: \figurenum{3}
917: \epsscale{0.9}
918: \plotone{fig3.ps}
919: \caption{}
920: \end{figure}
921: %
922: 
923: 
924: \end{document}
925: