1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2:
3: \documentclass[useAMS,usenatbib]{mn2e}
4: \usepackage{amssymb}
5: %\usepackage{amsmath}
6: \usepackage{graphics}
7: %\usepackage{Times}
8:
9: \topmargin -0.5in
10:
11: \def\TBD{\textbf{ToBeDone}}
12: \def\eqref#1{equation~(\ref{#1})}
13: \def\eqrefs#1#2{equations~(\ref{#1})-(\ref{#2})}
14: \def\apref#1{Appendix~\ref{#1}}
15: \def\scref#1{Section~\ref{#1}}
16: \def\binom#1#2{{#1 \choose #2}}
17: \def\arg{\mathop{\mathrm{arg}}\nolimits}
18: \def\sigmalevel#1{\ensuremath{#1}-\ensuremath{\sigma}}
19:
20: \newcommand{\obs}{\mathrm{\rm obs}}
21: \newcommand{\TD}{T_{\rm D}}
22:
23: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
24:
25: \title[Periastron Precession of TEPs]{Periastron Precession Measurements
26: in Transiting Extrasolar Planetary Systems at the Level of General Relativity}
27: \author[A. P\'al and B. Kocsis]{Andr\'as P\'al%
28: $^{1,2}$\thanks{E-mail: apal@cfa.harvard.edu}
29: and Bence Kocsis$^{1,3}$\thanks{E-mail: bkocsis@cfa.harvard.edu} \\
30: $^{1}$Harvard-Smithsonian Center for Astrophysics,
31: 60 Garden street,
32: Cambridge, MA, 02138, USA \\
33: $^{2}$Department of Astronomy, Lor\'and E\"otv\"os University,
34: P\'azm\'any P. st. 1/A,
35: Budapest H-1117, Hungary \\
36: $^{3}$Department of Atomic Physics, Institute of Physics, Lor\'and E\"otv\"os University,
37: P\'azm\'any P. st. 1/A,
38: Budapest H-1117, Hungary}
39: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
40:
41: \begin{document}
42:
43: \date{Accepted ..., Received ... ; in original form ...}
44:
45: \pagerange{\pageref{firstpage}--\pageref{lastpage}} \pubyear{2008}
46:
47: \maketitle
48:
49: \label{firstpage}
50:
51: \begin{abstract}
52: Transiting exoplanetary systems are surpassingly important among
53: the planetary systems since they provide the widest spectrum of
54: information for both the planet and the host star.
55: If a transiting planet is on an eccentric orbit,
56: the duration of transits $\TD$ is sensitive to
57: the orientation of the orbital ellipse relative to the line of sight.
58: The precession of the orbit results in a systematic variation
59: in both the duration of individual transit events and the
60: observed period between successive transits, $P_{\obs}$.
61: The periastron of the ellipse slowly precesses
62: due to general relativity and possibly the
63: presence of other planets in the system. This secular
64: precession can be detected through the long-term change in $P_{\obs}$
65: (transit timing variations, TTV) or in $\TD$ (transit duration variations, TDV).
66: We estimate the corresponding precession measurement precision for repeated
67: future observations of the
68: known eccentric transiting exoplanetary systems
69: (XO-3b, HD~147506b, GJ~436b and HD~17156b) using existing or planned
70: space-borne instruments. The TDV measurement improves the
71: precession detection sensitivity by orders of magnitude over the TTV measurement.
72: We find that TDV measurements over a $\sim4$\,year period can typically
73: detect the precession rate to a precision well exceeding the level predicted
74: by general relativity.
75: \end{abstract}
76:
77: \begin{keywords}
78: binaries: eclipsing -- planetary systems -- relativity -- methods: observational -- techniques: photometric
79: \end{keywords}
80:
81: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
82:
83: \section{Introduction}
84:
85: Since the discovery of the first transiting extrasolar planet
86: \citep{charbonneau2000,brown2001}, the number of such systems
87: has increased to more than 30\footnote{See http://exoplanet.eu
88: for up to date information}.
89: These transiting extrasolar planets (TEPs) provide unique
90: information on the properties of the system.
91: Based on the geometry provided by the transit light curve(s),
92: the inclination, the physical radius and mass, therefore the density
93: and the surface gravity can be derived, in addition to the mass of the planet.
94: Moreover, the time between successive transits can be measured with
95: an exceedingly high accuracy ($\sim 10^{-6}$~--~$10^{-7}$, relative to the period).
96: The detection of long--term transit timing variations can be used to learn more
97: beyond the properties of the parent-star system \citep{miralda2002,steffen2007}. They can be
98: indicative of the presence of other planetary companions \citep[see e.g.][]{holman2005,agol2005,millerricci2008},
99: co-orbital companions \citep[Trojans, see][]{ford2007},
100: or satellites \citep{simon2007} in the system, could provide information on the
101: oblateness of the host star, or can be used to detect the additional prograde periastron
102: precession predicted by general relativity (GR) \citep{miralda2002,heyl2007}.
103: Secular variations in the semimajor axis (and therefore in the transit timing)
104: are also predicted on the time scale of stellar life due to the anisotropic
105: light redistribution \citep[a.k.a. Yarkovski-effect, see][]{fabrycky2008}.
106: Furthermore, \citet{iorio2006} has shown that TEP observations can in principle
107: also test the gravitoelectric correction of GR
108: by measuring the radial velocity amplitude and transiting periodicity simultaneously,
109: in order to verify that the third Kepler's law requires a semimajor axis
110: dependent correction.
111:
112: In a pioneer study, \citet{miralda2002} derived the modification of the observed time period
113: between successive transits $P_{\obs}$, called transit timing variations (TTVs),
114: caused by the standard periastron precession due to GR \citep[e.g.][]{misner1973} and the perturbations of other planets if present.
115: Recently, \citet{heyl2007} have extended these studies and estimated the precision of
116: precession rate measurements for long--term mock observations of eccentric
117: transiting extrasolar planets (ETEPs). Both studies restricted to small eccentricities.
118: At that time, the existence of close eccentric planets was known
119: only through radial velocity measurements, and no ETEPs had been observed.
120: Since their publication, four transiting extrasolar planets have been discovered
121: with significant eccentricity: XO-3\lowercase{b} \citep{johnskrull2007},
122: HD~147506\lowercase{b} \citep[a.k.a. HAT-P-2,][]{bakos2007},
123: GJ~436\lowercase{b} \citep{gillon2007,butler2004},
124: and HD~17156\lowercase{b} \citep{fischer2007}.
125: Therefore it is now possible, for the first time, to make specific predictions
126: for future, long--term measurements of periastron precession effects for real
127: exoplanetary systems.
128:
129: In this paper we determine the precision by which repeated long--term future
130: ETEP observations will be able to
131: detect the periastron precession rate for existing systems.
132: In addition to TTVs, i.e. the
133: slow modulation of $P_{\obs}$ considered previously \citep{miralda2002,heyl2007},
134: the periastron precession also changes the time durations $\TD$ of individual transits.
135: We examine whether these transit duration variations (TDVs) can be used to
136: improve the sensitivity of periastron precession measurements.
137: We estimate the precession rate measurement precision for long term
138: repeated observations of $P_{\obs}$ and $\TD$ for the known ETEPs. Since several of the
139: observed ETEPs have large eccentricities, we derive expressions for both TTVs and TDVs
140: which are applicable for arbitrary eccentricities.
141: We estimate whether future observations of currently known ETEPs
142: will be able to reach the sensitivity necessary to test the prediction of GR,
143: using existing or planned space-borne instruments.
144: We refer the reader to a recent independent study by
145: \citet{jordan2008}, of precession rates in eccentric transiting
146: extrasolar planets.
147:
148: The next section of this paper introduces the geometrical description
149: which is the basis of our calculations, and derives the expected
150: transit timings and durations for planets orbiting a star with an
151: arbitrarily large eccentricity. In \S~\ref{s:realsystems}, we utilize our results for the
152: confirmed four ETEP systems, and give predictions
153: for future observations of periastron precession with space-borne observations.
154: Our conclusions are discussed in \S~\ref{s:summary}.
155:
156: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
157:
158: \section{Transit timings and durations for eccentric orbits}\label{s:transit}
159:
160: The reference frame used for the description of exoplanetary systems
161: as well as for binary/multiple stellar systems is fixed to the sky: the
162: plane of the sky is defined by the $(X+,Y+)$ while $Y+$ points towards
163: to north. For planetary transit observations, the line-of-sight lies
164: close to the orbital plane,
165: i.e. perpendicular to the plane of the sky.
166: The orbit is given by Cartesian coordinates $(\xi,\eta)$, where $\xi+$
167: is parallel to $X-$ and $\eta+$ oriented toward the observer (see also Fig.~\ref{fig:geometry}).
168: The Lagrange vector or eccentricity vector is given in these coordinates as
169: $(k,h)=(e\cos\omega,e\sin\omega)$.
170: Let us define the angle $\varphi_0$ as the angle relative to $\xi+$ in the orbital plane
171: at the instance\footnote{i.e. at the center of a transit} of the transit. From the definitions of $(\xi,\eta)$,
172: observing from Earth is equivalent to setting $\varphi_0=\pi/2$.
173:
174: \begin{figure}
175: \resizebox{8cm}{!}{\includegraphics{geometry.eps}}
176: \caption{%
177: The geometry of the orbit of the transiting planet. The plane of the
178: sky is defined by the $X+$ (west) and $Y+$ (north) axes while the $Z+$
179: axis points away from the Earth. The orbital plane is defined by the
180: axes $\xi+$ and $\eta+$, where $\xi+$ is anti-parallel with $X+$ axis,
181: $\eta+$ is in the plane of $(Y+,Z+)$ and the angle between $Y+$ and $\eta+$
182: is the inclination, nearly $90^{\circ}$. The major axis of the orbit
183: is marked by the dotted line.}\label{fig:geometry}
184: \end{figure}
185:
186: Now let us denote the mean longitude of the planet at the instance of the
187: transit by $\lambda$. For a circular orbit, $\lambda=\varphi_0$. From its standard
188: definition in celestial mechanics, it is straightforward to derive the mean longitude
189: for arbitrary eccentricities (see \apref{appendixlambdaderiv}). The result is
190: \begin{eqnarray}
191: \lambda & \equiv & \lambda(\varphi_0,k,h)\equiv \lambda(\varphi_0,e\cos\omega,e\sin\omega)= \nonumber \\
192: & = & \arg\left(k+\cos\varphi_0+\frac{he_{\perp}}{2-\ell},
193: h+\sin\varphi_0-\frac{ke_{\perp}}{2-\ell}\right) - \nonumber \\
194: & & -
195: \frac{e_{\perp}(1-\ell)}{1+e_{\parallel}}, \label{lambdaattransit}
196: \end{eqnarray}
197: where $e_{\parallel}=k\cos\varphi_0+h\sin\varphi_0$,
198: $e_{\perp}=k\sin\varphi_0-h\cos\varphi_0$ are the components of the eccentricity vector relative to the
199: line-of-sight, $\ell=1-\sqrt{1-e^2}$ is the oblateness of the orbit,
200: and $\arg(x,y)=\arctan(y/x)$ if $x\geq 0$ and $\pi+\arctan(y/x)$ otherwise.
201: Plugging in the observed values of $e$ and $\omega$,
202: \eqref{lambdaattransit} provides a simple way of calculating the mean longitude of the orbit
203: for an arbitrary transit observation. Note that this formalism omits the
204: direct usage of the mean anomaly, true anomaly and eccentric anomaly which
205: have no meanings for $e\to0$. All of our derived formulae are based
206: on the well--behaved parameters $\lambda$,
207: $k$, and $h$, and therefore are valid for arbitrary eccentricities.
208:
209: In the following subsections, we derive the expressions describing TTVs and TDVs, discuss the corresponding
210: observational implications and estimate the precision of periastron precession observations.
211:
212: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
213:
214: \begin{table*}
215: \caption{Basic data of the four known eccentric transiting
216: exoplanetary systems: the mass ($M_\star$) of the parent star,
217: period ($P$, in days), eccentricity ($e$)
218: and the argument of pericenter ($\omega$),
219: the half-duration of a transit event ($H$, in days)
220: and the impact parameter ($b$). In the last two columns, we provide
221: the calculated values of the secular period caused by the GR periastron precession, and the minimum mass -- semimajor axis ratio
222: for a hypothetical exterior perturber at $a_2\gtrsim 3a$ that would lead to the same magnitude of periastron precession.
223: }\label{table:basicdata}.
224: \begin{tabular}{llrlllllr}
225: \hline
226: System & $M_\star/M_{\odot}$ & $P$ (d) & $e$
227: & $\omega$ (degrees) & $H$ (d)
228: & $b$
229: & $P_{\rm sec}$ (years) & $\frac{m_2/M_{\oplus}}{(a_2/3a)^3}$\\
230: \hline
231: HD~147506b & $1.298\pm0.07$ & $ 5.63341 $ & $0.520\pm0.010$
232: & $179.3\pm3.6$ & $0.083$
233: & $0$
234: & $19790\pm740$ & $12.2$ \\
235:
236: XO-3b & $1.41\pm0.08$ & $ 3.19154$ & $0.260\pm0.017$
237: & $344.6\pm6.6$ & $0.050$
238: & $0.8$
239: & $9280\pm410$ & $15.9$ \\
240:
241: GJ~436b & $0.452\pm0.013$ & $ 2.64385 $ & $0.150\pm0.012$
242: & $351.0\pm1.2$ & $0.065$
243: & $0.85\pm0.02$
244: & $15180\pm400$ & $2.6$ \\
245:
246: HD~17156b & $1.2\pm0.1$ & $21.21725 $ & $0.6717\pm0.0027$
247: & $121.23\pm0.40$ & $0.098$
248: & $0.50\pm0.12$
249: & $143000\pm7900$ & $5.9$ \\
250: \hline
251: \end{tabular}
252: \end{table*}
253:
254: \subsection{Modulation of the transit period}
255:
256: The period between successive transits $P_{\obs}$ is modified by a possible slow precession of the orbital elements.
257: These modulations are referred to as transit timing variations (TTVs).
258: We derive the modulation of $P_{\rm obs}$ due to periastron precession in two steps.
259: First we demonstrate that the change in the period between successive transits is simply related to the change
260: in the mean longitude $\Delta \lambda$. Then relating the mean longitude shift to the periastron precession
261: rate we derive the expected TTV rate.
262:
263: Let $P_0$ be the orbital period, $n=2\pi/P_0$ be the mean angular
264: velocity. The mean longitude of the planet increases steadily in time,
265: $\lambda=nt+\lambda_0$. The variation in the mean longitude at
266: the transit center would result in
267: a variation in the transit cadence. During a transit at time $t_1$,
268: the mean longitude is $\lambda_{\rm tr}(t_1)=nt_1+\lambda_0$,
269: and after an observed revolution, at the instance $t_2=t_1+P_{\obs}$
270: it is $\lambda_{\rm tr}(t_2)+2\pi=nt_2+\lambda_0$.
271: Therefore the observed period between transits is
272: \begin{equation}\label{e:Pobsdef}
273: P_{\obs}=\frac{2\pi+\Delta \lambda}{n}=P_0\left(1+\frac{\Delta\lambda}{2\pi}\right)
274: \end{equation}
275: where $\Delta \lambda=\lambda_{\rm tr}(t_2)-\lambda_{\rm tr}(t_1)$.
276:
277: In the following we assume that the shift in the mean longitude is caused by the
278: perihelion shift $\Delta \omega = P_0 \dot\omega$
279: per period, (i.e. we assume a constant eccentricity), then from the chain rule
280: \begin{equation}\label{e:DLambda}
281: \Delta \lambda= \frac{\partial\lambda}{\partial \omega}\Delta\omega=
282: \frac{\partial\lambda}{\partial k} \frac{\partial k}{\partial \omega}\Delta\omega + \frac{\partial\lambda}{\partial h} \frac{\partial h}{\partial \omega}\Delta \omega.
283: \end{equation}
284: Here $\partial k/\partial \omega=-e\sin\omega$ and $\partial h/\partial \omega=e\cos\omega$, from definition (see above), and the partial derivatives $\partial\lambda/\partial k$ and
285: $\partial\lambda/\partial h$ can be found from
286: \eqref{lambdaattransit} and are given explicitly in the
287: Appendix (\ref{e:lambda_k}--\ref{e:lambda_h}). At transit, we get
288: \begin{equation}\label{e:lambda/omega}
289: \frac{\partial \lambda}{\partial \omega}=
290: \frac{e^2}{1+\sqrt{1-e^2}}+\sqrt{1-e^2}\frac{e^2+(2+e\sin\omega)e\sin\omega}{(1+e\sin\omega)^2}.
291: \end{equation}
292: Combining \eqrefs{e:Pobsdef}{e:lambda/omega} and defining
293: the secular period of the periastron precession as
294: $P_{\rm sec}=(2\pi)/\dot\omega$, we get
295: \begin{equation}
296: P_{\obs}=P_0+\frac{P_0^2}{2\pi} \frac{\partial\lambda}{\partial\omega} \dot\omega
297: =P_0+ \frac{\partial\lambda}{\partial\omega}\frac{P_0^2}{P_{\rm sec}}.\label{pobs1}
298: \end{equation}
299:
300: Since $\partial\lambda/\partial\omega$ itself is not constant due
301: to periastron precession,
302: the observed period between transits slowly changes. Differentiating \eqref{pobs1} with respect to time, we get
303: \begin{equation}\label{pobs1b}
304: \dot P_{\rm obs} = 2\pi\frac{\partial^2\lambda}{\partial\omega^2} \frac{P_0^2}{P_{\rm sec}^2}
305: =\frac{4\pi(1-e^2)^{3/2}e\cos\omega}{(1+e\sin\omega)^3}\frac{P_0^2}{P_{\rm sec}^2},
306: \end{equation}
307: since the partial derivative of \eqref{e:lambda/omega} with respect to $\omega$ is
308: \begin{equation}
309: \frac{\partial^2\lambda}{\partial\omega^2}=\frac{2(1-e^2)^{3/2}e\cos\omega}{(1+e\sin\omega)^3}. \label{d2ldo2}
310: \end{equation}
311: Note that this equation clearly shows that the small eccentricity
312: approximation $\partial^2\lambda/\partial\omega^2 \approx e\cos \omega$
313: \citep[e.g. used by][]{miralda2002,heyl2007} is very imprecise
314: for moderate to large eccentricites. Depending on the actual
315: value of $\omega$, \eqref{d2ldo2} can result even $6-8$ times
316: smaller or larger values for $\partial^2\lambda/\partial\omega^2$
317: as its first order approximation for eccentricities $0.5-0.7$.
318:
319: We have to mention here that the observed period and therefore the individual
320: transit timings are also affected by the light time effect (LTE). Since
321: the precession of an elliptical orbit causes the distance between
322: the host star and the planet at the transit instances to vary,
323: the light time delay will change for transit event to transit event.
324: The magnitude of this effect can be derived as follows.
325: The distance between the star and the planet at transit time
326: (i.e. when $\lambda=\varphi_0=\pi/2$) is
327: \begin{equation}
328: r=\frac{a(1-e^2)}{1+e\cos v}=\frac{a(1-e^2)}{1+e\cos(\varphi_0-\omega)}=
329: \frac{a(1-e^2)}{1+e\sin\omega}. \label{rattransit}
330: \end{equation}
331: This difference in the distance implies an additional $-r(1-\mu)/c$ time shift, respective to
332: the barycentric reference frame of the star--planet system. Here,
333: $\mu=M_{\rm p}/(M_{\rm p}+M_\star) \ll 1$ is the mass parameter, $a$
334: is the semimajor axis of the orbit and $c$
335: is the speed of light. Therefore, the correction in the observed period is
336: \begin{eqnarray}
337: P^{\rm +LTE}_{\obs} & = & P_{\obs}-\frac{a}{c}(1-e^2)\frac{\partial (1+e\sin\omega)^{-1}}{\partial\omega}P_0\dot\omega= \nonumber \\
338: & = & P_{\obs}+2\pi \frac{a}{c}\frac{(1-e^2)e\cos\omega}{(1+e\sin\omega)^2}\frac{P_0}{P_{\rm sec}},
339: \end{eqnarray}
340: where we neglected the barycentric correction.
341: Thus, the variation in this period (corrected for LTE) due to the
342: variation in $\omega$ is
343: \begin{equation}\label{e:LC}
344: \dot P^{\rm +LTE}_{\obs} = \dot P_{\obs} - 4\pi^2(1-e^2)\frac{e(e+e\cos^2\omega+\sin\omega)}{(1+e\sin\omega)^3}\frac{a}{c}\frac{P_0}{P_{\rm sec}^2}.
345: \end{equation}
346: Comparing \eqref{pobs1b} and (\ref{e:LC}), and assuming that
347: the motion of the planet is non-relativistic, $a/c\ll P_0$, we find that
348: $|\dot P^{\rm +LTE}_{\obs}-\dot P_{\obs}| \ll |\dot P_{\obs}|$.
349: We conclude that the period variation due to the LTE
350: is negligible compared to the period variation caused by the changing geometry.
351:
352: In summary, \eqrefs{pobs1}{pobs1b}, along with \eqref{e:lambda/omega}, give the
353: modulation of the actual observable period between transit events. These results are valid for arbitrary eccentricities
354: and are independent of the physical mechanism causing the secular precession of
355: the periastron. We calculate the secular precession period caused by general relativity \citep[see e.g.,][]{wald1984} using
356: \begin{equation}
357: P_{\rm sec} = \frac{c^2 (1-e^2)}{3(2\pi G M_\star)^{2/3}}
358: P_0^{5/3}, \label{pprecmmotion}
359: \end{equation}
360: where $M_\star$ is the mass of the parent star, $G$ is Newton's gravitational
361: constant. This secular period
362: is of order $10^4$--$10^5\,$years for the known ETEP systems
363: (see Table~\ref{table:basicdata} and \scref{s:realsystems}
364: for more details). We note that if other planets are also present in these systems, they might also cause additional periastron precession of a larger magnitude. The Yarkovski-effect \citep{fabrycky2008} and the tidal circularization \citep[see][and the references therein]{johnskrull2007} lead to negligible modifications for our purposes, as these effects are relevant on timescales exceeding $0.1\,$Gyr for these systems. In the following, we compare the precession measurement sensitivities with the general relativistic rate $P_{\rm sec}$ given by \eqref{pprecmmotion}.
365:
366:
367: \subsection{Modulation of the transit duration}
368:
369: Here we investigate how the periastron precession
370: affects the duration of a transit. Let us denote the half duration of the
371: transit by $H=\TD/2$, which we define as half the time between the instances when
372: the center of the planetary disk intersects the limb of the star, i.e. between
373: the center of the ingress and egress. Note that this
374: is not the time between the first and last contact.
375: This is important because the instances of the center of the ingress
376: and egress can be measured more accurately than their beginning or end.
377: Since we are interested in estimating the \emph{variations} of the duration
378: of the transit to leading order, we perform a first-order calculation,
379: i.e. assuming a constant apparent tangential velocity for the planet and neglecting
380: the changes in the impact parameter due to the elliptical orbit and/or
381: the curvature of the projection of the orbit due to the inclination. From
382: Kepler's Second Law and \eqref{rattransit}, one can calculate
383: the tangential distance $\Delta x$ traveled by the planet
384: during time $H$,
385: \begin{equation}
386: \Delta x=v_{\rm tan} H = an\frac{1+e\sin\omega}{\sqrt{1-e^2}} H
387: \end{equation}
388: (see \apref{appendixtangentvelocity} for the derivation of $v_{\rm tan}$).
389: This can be related to the impact parameter $b$ and
390: the radius of the star $R_\star$ for the geometry of
391: the transit as
392: \begin{equation}
393: \frac{\Delta x}{R_\star}= \sqrt{1-b^2}.
394: \end{equation}
395: Thus, to leading order,
396: \begin{equation}\label{e:H}
397: H = \frac{P_0}{2\pi}\left\{\frac{R_\star}{a}\frac{\sqrt{1-e^2}}{1+e\sin\omega}\sqrt{1-b^2} +
398: \mathcal{O}\left[\left(\frac{R_\star}{a}\right)^{3}\right]\right\}.
399: \end{equation}
400: Note that \eqref{e:H} depends on $\omega$ through the $(1+e\sin\omega)^{-1}$ term and
401: also implicitly through $b$,
402: \begin{equation}\label{e:b}
403: b=\frac{r\cos i}{R_\star} = \frac{a}{R_\star}\frac{1-e^2}{1+e\sin\omega}\cos i.
404: \end{equation}
405: The variation in $H$ caused by the variation in the periastron can be found from
406: \eqref{e:H} and (\ref{e:b}),
407: \begin{equation}
408: \frac{\partial H}{\partial\omega}=
409: \frac{e\cos\omega}{1+e\sin\omega}H\frac{1-2b^2}{1-b^2}. \label{dhdomega}
410: \end{equation}
411: Note that equation reflects the qualitative expections implied by
412: Kepler's Second Law. Namely, if an eccentric orbit advances, the
413: distance between the planet and the star will change. If this distance
414: decreases, the impact parameter will also decrease (resulting a longer
415: transit duration) but due to Kepler's Second Law, the apparent tangential
416: velocity of the transiting object will increase (resulting a shorter
417: transit duration). Therefore at a certain value of the inclination
418: and/or the impact parameter, the two effects cancel each other yielding
419: no TDV effect. Equation~\ref{dhdomega}
420: clearly shows that it occurs when the impact parameter is
421: $b=1/\sqrt{2}\approx0.707$.
422: The long-term variation in the duration of the transit is then
423: \begin{equation}\label{doth}
424: \dot H =\frac{\partial H}{\partial\omega}\dot\omega=\frac{\sqrt{1-e^2} e \cos \omega}{(1+e\sin \omega)^2}\frac{1-2b^2}{\sqrt{1-b^2}}\frac{R_{\star}}{a}\frac{P_0}{P_{\rm sec}}
425: \end{equation}
426: Comparing \eqref{pobs1b} and \eqref{doth} the TDV effect relates to the TTV effect as
427: \begin{equation}\label{e:TDV/TTV}
428: \frac{\dot H}{\dot P_{\rm obs}} =
429: %\frac{1+e\sin \omega}{1-e^2} \frac{R_{\star}}{4\pi a} \frac{P_{\rm sec}}{P_0} =
430: \frac{1+e\sin \omega}{6\pi}\frac{1-2b^2}{\sqrt{1-b^2}} \frac{R_{\star}}{R_{\rm Sch}}
431: \end{equation}
432: where $R_{\rm Sch}=2 G M/c^2$ is the Schwarzschild radius of the star. As an example, note that $R_{\star}/R_{\rm Sch}=2.5\times 10^{5}$ for the Sun.
433: Therefore, \eqref{e:TDV/TTV} shows that the TDV effect is always much larger than the TTV effect. In particular, the ratio is larger for increasing $b$. In the
434: limit $b\to 1$ \eqref{doth} breaks down because the periastron precession shifts the orbit out of the transiting region.
435:
436: \subsection{Observational implications}
437:
438: Now, using the results for the TDV and TTV effects, \eqref{pobs1b} and (\ref{doth}), we can estimate
439: how these timing and transit duration variations might be observed on
440: long timescales. In the following we discuss these observational implications.
441:
442: \subsubsection{Transit Timing Variations}
443: Equation~(\ref{pobs1b}) shows that the observed period between successive transits increases
444: or decreases at a practically constant rate during the observations, $\dot P_{\obs}$.
445: The time of the $m$th transit from an arbitrary epoch $T_0$ can be found from adding up the contributions of
446: the observed $m$ number of periods
447: \begin{eqnarray}
448: %T_n &=& T_0 + n P_{\obs}+\frac{n(n-1)}{2} \dot P_{\rm obs}P_0\\
449: T_m&=& T_0 + P_{\obs}m+Dm^2, \label{ttiming}
450: \end{eqnarray}
451: where $P_{\obs}\approx P_0$ denotes the time of the first observed orbit,
452: $D$ is the transit timing variation factor,
453: \begin{equation}
454: D=\frac{P_{0}\dot P_{\obs}}{2}=2\pi G_{\rm TV}(e,\omega)P_0\left(\frac{P_0}{P_{\rm sec}}\right)^2, \label{dtiming}
455: \end{equation}
456: where we have introduced the geometrical factor
457: \begin{equation}\label{e:G_TV}
458: G_{\rm TV}(e,\omega)=(1-e^2)^{3/2}\frac{e\cos\omega}{(1+e\sin\omega)^3}.
459: \end{equation}
460:
461: The chance of detecting the periastron precession increases
462: with the geometrical factor $G_{\rm TV}(e,\omega)$ which is related to
463: the alignment of the orbital ellipse with the line of sight.
464: The optimal value for detecting the precession is $\omega=0$ or $\pi$
465: for small eccentricities, i.e. the semimajor axis
466: should be perpendicular to the line of sight. For arbitrary eccentricities,
467: the optimal value for $\omega$ for the TTV observation is
468: \begin{equation}
469: \omega^{\rm best}_{\rm TV}=\frac{3}{2}\pi\pm\arccos\left(\frac{6e}{1+\sqrt{1+24e^2}}\right).
470: \end{equation}
471: which approaches $0^{\circ}$ and $180^{\circ}$ for small eccentricities as expected,
472: and $270^{\circ}$ for large eccentricities. In case of $e=0.5$,
473: $\omega_{\rm extr}=\{235.4^\circ,304.6^\circ\}$. The least favorable value for $\omega$ occurs
474: when $G_{\rm TV}(e,\omega)=0$, i.e. at $\omega^{\rm worst}_{\rm TV}=\{90^{\circ},270^{\circ}\}$,.
475:
476: \subsubsection{Transit Duration Variations}
477: Now let us turn to the TDV effect. The observed duration of the $m$th transit can be calculated in the same way,
478: namely
479: \begin{equation}
480: H_m=H_0+Fm,
481: \end{equation}
482: where $F$ is the shift in the transit duration per orbit. This factor is
483: \begin{equation}\label{e:F}
484: F= P_0 \dot H = 2\pi G_{\rm DV}(e,\omega,b)H\frac{P_0}{P_{\rm sec}},
485: \end{equation}
486: and
487: \begin{equation}\label{e:G_DV}
488: G_{\rm DV}(e,\omega,b)=\frac{1-2b^2}{1-b^2}\frac{e\cos\omega}{1+e\sin\omega}.
489: \end{equation}
490: The optimal orientation $\omega^{\rm best}_{\rm DV}$ for detecting the TDV effect for fixed $e$, $b$, and $P_0$ can be found by maximizing $ H G_{\rm DV}(e,\omega,b)$. The result is
491: \begin{equation}
492: \omega^{\rm best}_{\rm DV}=\frac{3}{2}\pi\pm\arccos\left(\frac{4e}{1+\sqrt{1+8e^2}}\right).
493: \end{equation}
494: and the worst orientation is at
495: $\omega^{\rm worst}_{\rm DV}=\{90^{\circ},270^{\circ}\}$,
496: just like for the TTV case. Comparing $\omega^{\rm best}_{\rm TV}$ and
497: $\omega^{\rm best}_{\rm DV}$ it is clear that the most favorable
498: orientation in terms of the two effects are similar, hence
499: the chance of detecting the periastron motion through
500: transit timing variations or transit duration variations
501: is correlated. Both effects go away if the eccentricity is oriented
502: parallel to the line of sight.
503: We also note that for moderate values of $e$ and small impact parameters,
504: $|G_{\rm TV}|\approx|G_{\rm DV}|$ which also implies that the
505: most favorable geometry for detecting either TTVs or TDVs is similar.
506:
507:
508: \subsection{Error analysis}
509: Next we estimate the parameter measurement
510: precision of the TTV and TDV effects for future observations.
511: We consider the repeated observation
512: of a particular transiting system over a total timespan $T_{\rm tot}$,
513: measuring the transit timing $T_m$ and duration $H_m$ for each transit
514: with respective errors $\sigma(T)$ and $\sigma(H)$. (We discuss the
515: specific values of $\sigma(T)$ and $\sigma(H)$ for transit observations
516: in Section~\ref{s:photometric}). For simplicity,
517: let us assume that these measurements are equidistant and in total
518: $N$ independent transits are observed, i.e. the $m$th transit is observed if
519: $m=0,d,\dots,(N-1)d$, where $d=T_{\rm tot}/(NP_0)$.
520:
521: \subsubsection{Transit Timing Variations}
522: Using \eqref{ttiming}, we can fit a second-order polynomial to these
523: observations with unknown coefficients $T_0$, $P_{\rm obs}$ and
524: $D$ by minimizing the merit function
525: \begin{equation}
526: \chi^2_{\rm TV} = \sum\limits_{m=0,d,\dots,(N-1)d}
527: \left[\frac{T_m - (T_0 + P_{\rm obs}m+Dm^2)}{\sigma(T)}\right]^2.
528: \end{equation}
529: The minimization of the above function results in a linear set of equations
530: in the parameters $p_{i}=\{T_0, P_{\rm obs}, D\}$. Assuming Gaussian errors,
531: the parameter estimation covariance matrix can be found from the
532: Fisher matrix method \citep{finn1992}:
533: \begin{equation}
534: \left< \delta p_i \delta p_j \right> = (\mathcal{F}^{-1})_{ij}
535: \end{equation}
536: Here $\mathcal{F}$ is the Fisher matrix defined as
537: \begin{equation}
538: \mathcal{F}_{ij} = \sum\limits_{m=0,d,\dots,(N-1)d} \frac{1}{\sigma^2(T)}\frac{\partial T^{\rm fid}_m}{\partial p_i} \frac{\partial T^{\rm fid}_m}{\partial p_j}
539: \end{equation}
540: where $T^{\rm fid}_m$ is the fiducial value of $T_m$ given by \eqref{ttiming}.
541: The marginalized expected squared parameter estimation error is given by the diagonal elements of the covariance error matrix $\sigma^2(p_i)=(\mathcal{F}^{-1})_{ii}$. In particular,
542: the resulting uncertainty of $D$ becomes
543: \begin{eqnarray}
544: \sigma(D) &=& \frac{\sqrt{180}\,\sigma(T)}{d^2 \sqrt{N(N^2-1)(N^2-4)}} \nonumber\\
545: &\approx& \sqrt{180} \left(\frac{P_0}{T_{\rm tot}}\right)^2 \left(1+\frac{5}{2N^2} \right) \frac{\sigma(T)}{\sqrt{N}}. \label{duncert}
546: \end{eqnarray}
547: Here, the first equality is valid for arbitrary $N$, while the second
548: is its first order approximation for large $N$. The leading order
549: approximation is verified against \citet{press1992}.
550:
551: \subsubsection{Transit Duration Variations}
552: We can repeat the same calculations as above for the observation of the
553: half transit duration $H_m$ to measure the variation factor $F$.
554: The merit function in this case
555: \begin{equation}
556: \chi^2_{\rm DV} = \sum\limits_{m=0,d,\dots,(N-1)d}
557: \left[\frac{H_m - (H_0 + Fm)}{\sigma(H)}\right]^2,
558: \end{equation}
559: has to be minimized for the same set of observations.
560: This minimization again leads to a linear set of equations in the
561: parameters $H_0$ and $F$. The Fisher matrix method in this case
562: gives the uncertainty in $F$ as
563: \begin{eqnarray}
564: \sigma(F) &=& \frac{\sqrt{12}\sigma(H)}{d\sqrt{N(N^2-1)}}\approx \nonumber\\
565: &\approx& \sqrt{12} \frac{P_0}{T_{\rm tot}} \left(1+\frac{1}{2N^2} \right)
566: \frac{\sigma(H)}{\sqrt{N}}. \label{funcert}
567: \end{eqnarray}
568:
569: \begin{figure}
570: \resizebox{8cm}{!}{\includegraphics{tdvksig.eps}}
571: \caption{%
572: The significance of detecting the GR periastron precession $|F_0|/\sigma_0(F)$
573: through the TDV effect as a function
574: of orbital period and eccentricity. The transit duration is assumed to be measured for $4$~years each with a $5$~sec error,
575: for a Sun-like star on a non-inclined orbit.
576: Increasing the mass or radius of the star, or the impact parameter increases the
577: detection significance (see text).
578: }\label{fig:tdvksig}
579: \end{figure}
580: Figure~\ref{fig:tdvksig} shows the detection significance of the TDV
581: measurement $|F|/\sigma(F)$, if the precession rate in $F$ is given by
582: the general relativistic formula, \eqref{e:F}.
583: Here each transit is assumed to be measured (i.e. $d=1$) with a precision $\sigma(H)=5$\,sec
584: for a total observation time of $T_{\rm tot}=4$\,years. These assumptions are
585: realistic for the future Kepler mission (see \S~\ref{s:photometric} below).
586: Other parameters are $M_{\star}=M_{\odot}$, $R_{\star}=R_{\odot}$, $b=0$, and
587: we averaged over the possible orientations of $\omega$.
588: For other parameters,
589: \begin{eqnarray}
590: \frac{|F|/\sigma(F)}{|F_{0}|/\sigma(F_{0})} & = & \frac{1}{d}\frac{1}{\sqrt{1-b^2}}
591: \left(\frac{R_{\star}}{R_{\odot}}\right) \left(\frac{M_{\star}}{M_{\odot}}\right)^{1/3}\cdot \nonumber \\
592: & & \cdot
593: \left(\frac{\sigma(H)}{5\,{\rm sec}}\right)^{-1}
594: \left(\frac{T_{\rm tot}}{4\,{\rm yr}}\right)^{3/2},
595: \end{eqnarray}
596: implying that the detection significance can be even better.
597:
598: Figure~\ref{fig:tdvksig} clearly shows that the chances of detecting the
599: precession effects
600: through the TDV effect is encouraging. The detection significance of the
601: general relativistic precession of a transiting
602: exoplanet with eccentricity $e\gtrsim 0.2$ and period $P\lesssim 5$~days, is
603: typically over the \sigmalevel{1} level.
604: Generally, \eqref{duncert} and \eqref{funcert} can be used directly to
605: check what kind of observations are required to detect
606: the precession of the periastron through the TTV or TDV methods, respectively.
607:
608:
609: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
610:
611: \section{The case of XO-3\lowercase{b},
612: HD~147506\lowercase{b}, GJ436\lowercase{b} and HD~17156\lowercase{b}}
613: \label{s:realsystems}
614:
615: As of this writing, four TEPs are known with a non-zero eccentricity within
616: \sigmalevel{3}, namely
617: XO-3\lowercase{b} \citep{johnskrull2007},
618: HD~147506\lowercase{b} \citep{bakos2007},
619: GJ~436b\lowercase{b} \citep{butler2004,gillon2007},
620: and HD~17156b\lowercase{b} \citep{fischer2007,barbieri2007}.
621: The planet TrES-1 \citep{alonso2004} has also been reported as an object with
622: non-zero eccentricity, however, it is zero within \sigmalevel{2} thus we
623: omit from our analysis. We note here that recently both GJ~436
624: and HD~17156 have been suggested to have another planetary companions
625: \citep[see][]{ribas2008,short2008}.
626: The secular period of the periastron motion
627: are determined by the mass of the star $M_\star$, the orbital period $P_0$,
628: and the eccentricity $e$,
629: while the timing variation constant $D$ is also affected by the
630: actual argument of pericenter, $\omega$. The transit duration variation
631: factor $F$ is affected indirectly by the geometrical ratio
632: $a/R_\star$ and directly by the impact parameter $b$.
633: These parameters are summarized in the first seven columns of Table~\ref{table:basicdata}
634: for these four ETEP systems. The derived GR periastron precession period,
635: $P_{\rm sec}$ can be found in the 8th column of the table.
636:
637: In addition to the inevitable periastron precession caused by GR, there
638: might be other sources of perturbations causing periastron precession.
639: The last column gives the minimum mass to semimajor axis ratio
640: of a hypothetical exterior perturber (e.g. a planet or an asteroid belt), in Earth mass units, which
641: causes the same periastron presession rate as that caused by the general
642: relativity. This estimate based on \citet{price1979}, and is valid for
643: $a_2\gtrsim 3a$ and for non-resonant cases.
644: Note that the minimum mass of the perturber scales with the third power of the
645: semimajor axis ratio to cause a comparable precession rate as GR.
646: The numbers show that the precession caused by additional planets
647: in the system, if present, can easily cause a larger precession rate than GR.
648: In case of orbital resonances with exterior planets,
649: the precession rate can be even larger \citep{holman2005,agol2005}.
650: To be conservative, we examine whether the precession rate can be measured
651: to a precision better than that corresponding to GR.
652:
653: To obtain a high significance, \sigmalevel{k} detection of the
654: periastron precession using
655: transit timing variations, we need $k\sigma(D)\approx|D|$.
656: Using \eqref{duncert},
657: the total number of transits necessary to measure $D$ with this precision is
658: \begin{equation}
659: N_{\rm TV}=180\left[\frac{\sigma(T)}{k|D|}\right]^2\left(\frac{P_0}{T_{\rm tot}}\right)^4.
660: \end{equation}
661: Thus, the number of such required transit observations
662: is extremely sensitive to the orbital period $P_0$ and the observation timespan.
663: Table~\ref{table:ttvdata} gives the corresponding values of
664: the transit timing variation factors for the known ETEP systems
665: and this number of observations, assuming a $T_{\rm tot}=20$~year long
666: observational timespan, timing precision of $\sigma(T)=2~{\rm sec}$
667: and \sigmalevel{3} sensitivity of the GR periastron precession
668: level. The table shows the recently discovered XO-3b system is a promising
669: candidate to detect the GR periastron precession through the TTV effect,
670: while the other ETEP systems require unrealistically many observations. We note
671: that if other perturbing planets are present in these systems and lead to a
672: precession rate that is larger by a factor of 10 than the GR precession rate,
673: then the number of detections (during the same $T_{\rm tot}$ timespan) is
674: lower by a factor of 100. It is interesting that the best candidate (by far) is
675: XO-3b even though its eccentricity is not as large as that of
676: HD~147506b or HD~17156b.
677:
678: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
679:
680: \begin{table}
681: \caption{Transit timing variation factor ($D$, in seconds) and the
682: number of transits
683: ($N_{20\,{\rm y},\sigmalevel{3},2\,{\rm sec}}$) what should be
684: detected almost uniformly in a 20 year long timespan, each with an error of
685: $2$ second to confirm the precession within \sigmalevel{3}.
686: }\label{table:ttvdata}
687: \begin{center}\begin{tabular}{llr}
688: \hline
689: System & $D$ (seconds)
690: & $N_{20\,{\rm y},~\sigmalevel{3},~2\,{\rm sec}}$ \\
691: \hline
692: HD~147506b & $-(5.9\pm0.7)\cdot10^{-7}$
693: & $6600$ \\
694: XO-3b & $+(4.3\pm0.6)\cdot10^{-7}$
695: & $1280$ \\
696: GJ~436b & $+(5.0\pm0.5)\cdot10^{-8}$
697: & $44160$ \\
698: HD~17156b & $-(6.9\pm0.8)\cdot10^{-8}$
699: & $97\cdot10^{6}$ \\
700: \hline
701: \end{tabular}\end{center}
702: \end{table}
703:
704: \begin{table}
705: \caption{Transit duration variation factor ($F$, in seconds) and the
706: number of transits
707: ($N_{2\,{\rm y},~\sigmalevel{3},~2\,{\rm sec}}$) what should be
708: detected almost uniformly in a 4 year long timespan, each with an error of
709: $2$\,sec to confirm the precession within \sigmalevel{3}.
710: }\label{table:tdvdata}
711: \begin{center}\begin{tabular}{llr}
712: \hline
713: System & $F$ (seconds)
714: & $N_{4\,{\rm y},~\sigmalevel{3},~2\,{\rm sec}}$ \\
715: \hline
716: HD~147506b & $-(1.8\pm0.3)\cdot10^{-2}$
717: & $20$ \\
718: XO-3b & $-(5.4\pm0.6)\cdot10^{-3}$
719: & $70$ \\
720: GJ~436b & $-(4.1\pm0.5)\cdot10^{-3}$
721: & $85$ \\
722: HD~17156b & $-(3.2\pm0.2)\cdot10^{-3}$
723: & $8970$ \\
724: \hline
725: \end{tabular}\end{center}
726: \end{table}
727:
728: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
729:
730: Let us now turn to the observational constraints for the detection
731: of transit duration variations. Using \eqref{funcert},
732: the total number of required observations within $T_{\rm tot}$ is
733: \begin{equation}
734: N_{\rm DV}=12\left[\frac{\sigma(H)}{k|F|}\right]^2
735: \left(\frac{P_0}{T_{\rm tot}}\right)^2.
736: \end{equation}
737: Note that $N_{\rm DV}$ is not as sensitive to the $P_0/T_{\rm tot}$
738: ratio as $N_{\rm TV}$, and implies that a smaller number of observations is
739: typically necessary.
740:
741: In Table~\ref{table:tdvdata} we present the values of the transit
742: duration variation factor, $F$, and the number of required observations
743: to reach the same \sigmalevel{3} confidence for detecting the GR
744: periastron precession\footnote{Note that since the measurement of the
745: TDV effect relies on fitting 2 parameters, instead of 3 parameters
746: for the TTV effect, the \sigmalevel{3} confidence corresponds to a higher
747: confidence level for the TDV effect.}.
748: Here we assumed a shorter observation timespan,
749: $T_{\rm tot}=4$~year (i.e. shorter compared to the 20~year long timespan
750: necessary for the detection of the TTV effect),
751: the same timing precision
752: of $\sigma(H)=2~{\rm sec}$ and the same level of detection, \sigmalevel{3}.
753: The best known ETEP system for TDV detection is therefore HD~147506b,
754: but the number of necessary observations
755: is feasible for XO-3b and GJ~436b as well.
756:
757: \subsection{Photometric detection}
758: \label{s:photometric}
759:
760: The precision for measuring the transit timing and transit duration
761: for a photometric observation can be estimated as follows. Since
762: the time of the ingress ($T_{\rm I}$) and the time of the egress
763: ($T_{\rm E}$) -- i.e. when the center of the planet crosses the
764: limb of the star inwards or outwards, respectively --
765: defines both the time of the transit center and
766: the half duration like
767: \begin{eqnarray}
768: T & = & \frac12(T_{\rm E}+T_{\rm I}), \\
769: H & = & \frac12(T_{\rm E}-T_{\rm I}),
770: \end{eqnarray}
771: moreover $T_{\rm I}$ and $T_{\rm E}$ can be treated as uncorrelated variables,
772: therefore the uncertainties of the transit time and half duration would
773: be nearly the same, i.e. $\sigma(T)\approx\sigma(H)$. We have estimated
774: these uncertainties for the four distinct planets
775: using Monte-Carlo simulations by fitting
776: transit light curves on mock data sets. We have used the
777: observed planetary parameters as an input for these artificial light curves.
778: The fit was performed assuming
779: quadratic limb darkening \citep[see][]{mandel2002} in the Sloan $z'$ band.
780: The mock light curves were sampled with $\Delta\tau_1=1\,{\rm sec}$
781: cadence and an additional Gaussian noise of $\sigma_1(m)=1\,{\rm mmag}$
782: was added. The resulting uncertainties, $\sigma_{1,1}(T)$ and $\sigma_{1,1}(H)$
783: for the four planets are presented in Table~\ref{table:ttvtdvuncert}. Since
784: the depth of the four transits are nearly the same (see the appropriate
785: normalized radii, $p=R_p/R_\star$, all between $0.068\lesssim p\lesssim0.085$),
786: the uncertainties $\sigma_{1,1}(T)$ and $\sigma_{1,1}(H)$ are almost the same
787: for the four cases. Using these normalized values, one can
788: easily estimate the uncertainties for arbitrary sampling cadence
789: $\Delta \tau$ and photometric precision $\sigma(m)$ using
790: \begin{eqnarray}
791: \sigma(T) & \approx & \sigma_{1,1}(T)\frac{\sigma(m)}{1\,\rm mmag}
792: \sqrt{\frac{\Delta\tau}{1\,\rm sec}}, \\
793: \sigma(H) & \approx & \sigma_{1,1}(H)\frac{\sigma(m)}{1\,\rm mmag}
794: \sqrt{\frac{\Delta\tau}{1\,\rm sec}}.
795: \end{eqnarray}
796: For comparison, note that the expected photometric precision of the Kepler
797: space telescope \citep[see][]{borucki2007} is $1\,{\rm mmag}$
798: for observing a light curve of a bright, $M_v=8.8$ star with a
799: $1\,{\rm sec}$ sampling cadence. Since
800: the star XO-3 has almost the same apparent magnitude, it is clear,
801: the transit durations
802: would be detected with an accuracy of $\sigma(H)\approx 5\,{\rm sec}$
803: if this star was in the field of Kepler.
804: Therefore, \eqref{funcert} and Table~\ref{table:tdvdata} shows
805: that the transit duration variations
806: would be detectable for HD~147506(b) or XO-3(b)--like systems
807: due to the GR periastron precession,
808: within \sigmalevel{3} confidence
809: with a Kepler--type mission within approximately 3 or 4 years,
810: respectively.
811:
812:
813: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
814:
815: \begin{table}
816: \caption{Uncertainties of the transit time and transit duration measurements
817: for the four known ETEPs, assuming Sloan $z'$-band photometric data
818: taken with a 1\,sec cadence and 1\,mmag photometric precision.
819: }\label{table:ttvtdvuncert}
820: \begin{center}\begin{tabular}{lll}
821: \hline
822: System & $\sigma_{1,1}(T)$ (sec)
823: & $\sigma_{1,1}(H)$ (sec) \\
824: \hline
825: HD~147506b & 5.3 & 4.8 \\
826: XO-3b & 6.9 & 4.7 \\
827: GJ~436b & 6.7 & 8.4 \\
828: HD~17156b & 5.4 & 6.1 \\
829: \hline
830: \end{tabular}\end{center}
831: \end{table}
832:
833: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
834:
835:
836:
837: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
838:
839: \section{Summary}
840: \label{s:summary}
841:
842: The first four eccentric transiting exoplanetary
843: systems have been discovered during 2007. The precession of an eccentric orbit
844: causes variations both in the transit timings and transit durations.
845: We estimated the significance of measuring the corresponding observable effects
846: compared to the inevitable precession rate of general relativity.
847: We applied these
848: calculations to predict the significance of measuring the
849: effect for the four known eccentric transiting planetary systems.
850: Our calculations show that a space-borne telescope is
851: adequate to detect the change in the transit durations
852: to a high significance better than the GR periastron precession rate
853: within a 3~--~4 year timespan (in a continuous observing mode).
854: The same kind of instruments would need more than a decade
855: to detect the corresponding transit time variations to this sensitivity
856: even for the most optimistic known system.
857:
858: The CoRoT mission has already found two transiting planets
859: \citep[see][]{barge2008,alonso2008} and there are two known
860: planets in the planned field-of-view of the Kepler mission
861: \citep[see][]{odonovan2006,pal2008}.
862: Our results suggest that if an \emph{eccentric} transiting planet
863: is found in the Kepler or CoRoT field, these missions
864: will be able to measure the periastron
865: precession rate to a very high significance within their mission lifetime
866: or with the support of ground-based observations on a longer time scale.
867: This will provide an independent test of the theory of
868: general relativity and will also be useful for testing for the
869: presence of other planets in these systems.
870:
871: \section*{Acknowledgments}
872:
873: The authors would like to thank the hospitality and support of
874: the Harvard-Smithsonian Center for Astrophysics where this
875: work was partially carried out.
876: We thank Andres Jordan for useful comments on the manuscript.
877: BK acknowledges support from OTKA grant No.~68228.
878:
879: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
880:
881: \begin{thebibliography}{99}
882:
883: %On detecting terrestrial planets with timing of giant planet transits
884: \bibitem[\protect\citeauthoryear{Agol et al.}{2005}]{agol2005}
885: Agol, E., Steffen, J., Sari, R., \& Clarkson, W.,
886: 2005, MNRAS, 359, 567
887:
888: % TrES-1:
889: \bibitem[\protect\citeauthoryear{Alonso et al.}{2004}]{alonso2004}
890: Alonso, R. et al.,
891: 2004, ApJ, 613, 153
892:
893: % Corot-exo-II
894: \bibitem[\protect\citeauthoryear{Alonso et al.}{2008}]{alonso2008}
895: Alonso, R. et al.,
896: 2008, astro-ph:0803.3207
897:
898: % HAT-P-2:
899: \bibitem[\protect\citeauthoryear{Bakos et al.}{2007}]{bakos2007}
900: %Bakos, G. \'A., Kov\'acs, G., Torres, G., Fischer, D. A., Latham, D. W., Noyes, R. W., Sasselov, D. D., Mazeh, T., Shporer, A., Butler, R. P., Stefanik, R. P., Fernandez, J. M., Sozzetti, A., P\'al, A., Johnson, J., Marcy, G. W., Winn, J., Sip\H{o}cz, B., L\'az\'ar, J., Papp, I. \& S\'ari, P.,
901: Bakos, G. \'A. et al.,
902: 2007, ApJ, 670, 826
903:
904: % transits of HD 17156b:
905: \bibitem[\protect\citeauthoryear{Barbieri et al.}{2007}]{barbieri2007}
906: %Barbieri, M., Alonso, R., Laughlin, G., Almenara, J. M., Bissinger, R., Davies, D., Gasparri, D., Guido, E., Lopresti, C., Manzini, F. \& Sostero, G.,
907: Barbieri, M. et al.,
908: 2007, A\&A, 476, 13
909:
910: % Corot-exo-1
911: \bibitem[\protect\citeauthoryear{Barge et al.}{2008}]{barge2008}
912: Barge, P. et al.,
913: 2008, astro-ph:0803.3202
914:
915: % KEPLER mission status:
916: \bibitem[\protect\citeauthoryear{Borucki et al.}{2007}]{borucki2007}
917: Borucki, W. J.~et al.,
918: 2007, ASP Conf.~Ser., 366, 309
919:
920: % HST, HD209458
921: \bibitem[\protect\citeauthoryear{Brown et al.}{2001}]{brown2001}
922: %Brown, T. M., Charbonneau, D., Gilliland, R. L., Noyes, R. W. \& Burrows, A.,
923: Brown, T. M. et al.,
924: 2001, ApJ, 552, 699
925:
926: % GJ 436:
927: \bibitem[\protect\citeauthoryear{Butler et al.}{2004}]{butler2004}
928: %Butler, R. Paul; Vogt, Steven S.; Marcy, Geoffrey W.; Fischer, Debra A.; Wright, Jason T.; Henry, Gregory W.; Laughlin, Greg; Lissauer, Jack J.
929: Butler et al.,
930: 2004, ApJ, 617, 580
931:
932: % first transiting planet: HD209458:
933: \bibitem[\protect\citeauthoryear{Charbonneau et al.}{2000}]{charbonneau2000}
934: Charbonneau, D., Brown, T. M., Latham, D. W. \& Major, M.,
935: 2000, ApJ, 529, 45
936:
937: % HD17156:
938: \bibitem[\protect\citeauthoryear{Fischer et al.}{2007}]{fischer2007}
939: %Fischer, D. A., Vogt, S. S., Marcy, G. W.; Butler, R. P., Sato, B., Henry, G. W., Robinson, S., Laughlin, G., Ida, S., Toyota, E., Omiya, M., Driscoll, P., Takeda, G., Wright, J. T. \& Johnson, J. A.,
940: Fischer, D. A. et al.,
941: 2007, ApJ, 669, 1336
942:
943: % Radiative trusters:
944: \bibitem[\protect\citeauthoryear{Fabrycky}{2008}]{fabrycky2008}
945: % Daniel Fabrycky
946: Fabrycky, D.,
947: 2008, astro-ph:0803.1839
948:
949: %Fisher matrix:
950: \bibitem[\protect\citeauthoryear{Finn}{1992}]{finn1992}
951: % Finn, L. S.,
952: Finn, L. S.,
953: 1992, Phys.~Rev.~D, 46, 5236
954:
955: % TTV, Trojans:
956: \bibitem[\protect\citeauthoryear{Ford \& Holman}{2007}]{ford2007}
957: Ford, E. B. \& Holman, M. J.,
958: 2007, ApJ, 664, 51
959:
960: % transits of GJ 436:
961: \bibitem[\protect\citeauthoryear{Gillon et al.}{2007}]{gillon2007}
962: %Gillon, M.; Pont, F.; Demory, B.-O.; Mallmann, F.; Mayor, M.; Mazeh, T.; Queloz, D.; Shporer, A.; Udry, S.; Vuissoz, C.
963: Gillon, M. et al.,
964: 2007, A\&A, 472, 13
965:
966: % Using long-term transit timing to detect terrestrial planets:
967: \bibitem[\protect\citeauthoryear{Heyl \& Gladman}{2007}]{heyl2007}
968: % Heyl, Jeremy S.; Gladman, Brett J.
969: Heyl, J. S. \& Gladman, B. J.,
970: 2007, MNRAS, 377, 1511
971:
972: %The Use of Transit Timing to Detect Terrestrial-Mass Extrasolar Planets
973: \bibitem[\protect\citeauthoryear{Holman \& Murray}{2005}]{holman2005}
974: Holman, M. J, \& Murray, N. W.,
975: 2005, Science, 307, 1288
976:
977:
978: % GR: RV + ttv combinations
979: \bibitem[\protect\citeauthoryear{Iorio}{2006}]{iorio2006}
980: Iorio, L.,
981: 2006, NewA, 11, 490
982:
983: % Jordan & Bakos
984: \bibitem[\protect\citeauthoryear{Jordan \& Bakos}{2008}]{jordan2008}
985: Jordan,A. \& Bakos, G,
986: 2008, ApJ, in press
987:
988: % XO-3b:
989: \bibitem[\protect\citeauthoryear{Johns-Krull et al.}{2007}]{johnskrull2007}
990: %Johns-Krull, C .M., McCullough, P. R., Burke, C. J., Valenti, J. A., Janes, K. A., Heasley, J. N., Prato, L., Bissinger, R., Fleenor, M., Foote, C. N., Garcia-Melendo, E., Gary, B. L., Howell, P. J., Mallia, F., Masi, G. \& Vanmunster, T.,
991: Johns-Krull, C .M. et al.,
992: 2007, astro-ph:0712.4283
993:
994: % TLC modelling:
995: \bibitem[\protect\citeauthoryear{Mandel \& Agol}{2002}]{mandel2002}
996: Mandel, K., Agol, E., 2002,
997: ApJ, 580, 171
998:
999: % HD209458, MOST, TTVs:
1000: \bibitem[\protect\citeauthoryear{Miller-Ricci et al.}{2008}]{millerricci2008}
1001: Miller-Ricci, E. et al.,
1002: 2008, astro-ph:0802.0718
1003:
1004: % possible TTVs
1005: \bibitem[\protect\citeauthoryear{Miralda-Escude}{2002}]{miralda2002}
1006: Miralda-Escude, J.,
1007: 2002, ApJ, 564, 1019
1008:
1009: % Gravitation:
1010: \bibitem[\protect\citeauthoryear{Misner, Thorne, \& Wheeler}{1973}]{misner1973}
1011: Misner, C. W., Thorne, K., \& Wheeler, J. A.,
1012: 1973,
1013: Gravitation,
1014: Second Edition, W. H. Freeman and Company, San Francisco
1015:
1016: % In the kepler field: TrES-2:
1017: \bibitem[\protect\citeauthoryear{O'Donovan et al.}{2006}]{odonovan2006}
1018: % O'Donovan, Francis T.; Charbonneau, David; Mandushev, Georgi; Dunham, Edward W.; Latham, David W.; Torres, Guillermo; Sozzetti, Alessandro; Brown, Timothy M.; Trauger, John T.; Belmonte, Juan A.; Rabus, Markus; Almenara, Jos M.; Alonso, Roi; Deeg, Hans J.; Esquerdo, Gilbert A.; Falco, Emilio E.; Hillenbrand, Lynne A.; Roussanova, Anna; Stefanik, Robert P.; Winn, Joshua N.
1019: O'Donovan, F. T. et al.,
1020: 2006, ApJ, 651, 61
1021:
1022: % In the Kepler field: HAT-P-7:
1023: \bibitem[\protect\citeauthoryear{P\'al et al.}{2008}]{pal2008}
1024: % Pal, A.; Bakos, G. A.; Torres, G.; Noyes, R. W.; Latham, D. W.; Kovacs, Geza; Marcy, G. W.; Fischer, D. A.; Butler, R. P.; Sasselov, D. D.; Sipocz, B.; Esquerdo, G. A.; Kovacs, Gabor; Stefanik, R.; Lazar, J.; Papp, I.; Sari, P.
1025: P\'al, A. et al.,
1026: 2008, astro-ph:0803.0746
1027:
1028: % Num recipes:
1029: \bibitem[\protect\citeauthoryear{Press et al.}{1992}]{press1992}
1030: Press, W. H., Teukolsky, S. A., Vetterling, W. T. \& Flannery, B. P.,
1031: 1992,
1032: Numerical Recipes in C: the art of scientific computing,
1033: Second Edition, Cambridge University Press
1034:
1035: % Periastron precession caused by an additional exterior planet
1036: \bibitem[\protect\citeauthoryear{Price \& Rush}{1979}]{price1979}
1037: Price, M. P., \& Rush, W. E.,
1038: 1979, Am. J. Phys., 47, 531
1039:
1040: % GJ 436:
1041: \bibitem[\protect\citeauthoryear{Ribas et al.}{2008}]{ribas2008}
1042: Ribas, I., Font-Ribera, A. \& Beaulieu, J.,
1043: 2008, ApJ, 677, 59
1044:
1045: % HD17156, other planet:
1046: \bibitem[\protect\citeauthoryear{Short et al.}{2008}]{short2008}
1047: Short, D., Welsh, W. F., Orosz, J. A. \& Windmiller G.,
1048: 2008, astro-ph:0803.2935
1049:
1050: % Exomoons:
1051: \bibitem[\protect\citeauthoryear{Simon et al.}{2007}]{simon2007}
1052: Simon, A., Szatm\'ary, K. \& Szab\'o, Gy. M.,
1053: 2007, A\&A, 470, 727
1054:
1055: % TTV issues:
1056: \bibitem[\protect\citeauthoryear{Steffen \& Agol}{2007}]{steffen2007}
1057: Steffen, J. H. \& Agol, E.,
1058: 2007, ASP Conf.~Ser, 366, 158
1059:
1060: % GR:
1061: \bibitem[\protect\citeauthoryear{Wald}{1984}]{wald1984}
1062: Wald, R. M.,
1063: 1984
1064: General Relativity,
1065: The University of Chicago Press
1066:
1067: \end{thebibliography}{}
1068:
1069: % This is the "This paper has been typeset from a TeX/LaTex file ..." trailer.
1070: % \bsp
1071:
1072: \appendix
1073: \section{Mean longitude at the transit instances}
1074: \label{appendixlambdaderiv}
1075:
1076: The derivation of \eqref{lambdaattransit} goes as follows. According to
1077: Kepler's equation, $E-e\sin E=M=\lambda-\omega$, one can write
1078: $\lambda=\omega+E-e\sin E$. The only thing what is to be done is to
1079: calculate the eccentric anomaly $E$ for the instance when the orbiting
1080: body intersect the semi-line with the argument angle $\varphi_0$. The
1081: latter means that the true anomaly $v$ of the body is $v=\varphi_0-\omega$,
1082: by definition. The relation between the eccentric and true anomaly
1083: is
1084: \begin{equation}
1085: \tan\frac{E}{2}=\sqrt{\frac{1-e}{1+e}}\tan{\frac{v}{2}},
1086: \end{equation}
1087: which is equivalent with
1088: \begin{eqnarray}
1089: \cos E & = & \frac{e+\cos v}{1+e\cos v}, \label{coseatt}\\
1090: \sin E & = & \frac{\sqrt{1-e^2}\sin v}{1+e\cos v}. \label{sineatt}
1091: \end{eqnarray}
1092: Using the addition theorem, the sine and cosine of the angle $\omega+E$
1093: can be written as:
1094: \begin{eqnarray}
1095: \cos(\omega+E) & = & \cos\omega\frac{e+\cos(\varphi_0-\omega)}{1+e\cos(\varphi_0-\omega)} - \nonumber \\
1096: & & - \sin\omega\frac{\sqrt{1-e^2}\sin(\varphi_0-\omega)}{1+e\cos(\varphi_0-\omega)}, \\
1097: \sin(\omega+E) & = & \sin\omega\frac{e+\cos(\varphi_0-\omega)}{1+e\cos(\varphi_0-\omega)} + \nonumber \\
1098: & & + \cos\omega\frac{\sqrt{1-e^2}\sin(\varphi_0-\omega)}{1+e\cos(\varphi_0-\omega)}.
1099: \end{eqnarray}
1100: Thus, the mean longitude itself is going to be
1101: \begin{eqnarray}
1102: \lambda & = & \omega+E-e\sin E=\arg\left[\cos(\omega+E),\sin(\omega+E)\right]- \nonumber \\
1103: & & - e\frac{\sqrt{1-e^2}\sin v}{1+e\cos v}.
1104: \end{eqnarray}
1105: If both arguments of the above $\arg[\cdot,\cdot]$ function is
1106: multiplied by the always positive common denominator
1107: $1+e\cos(\varphi_0-\omega)$, one gets after some simplification:
1108: \begin{eqnarray}
1109: \omega+E & = & \arg\left[k+\cos\varphi_0+\frac{h(k\sin\varphi_0-h\cos\varphi_0)}{1+\sqrt{1-e^2}}, \right. \nonumber \\
1110: & & \left.h+\sin\varphi_0-\frac{k(k\sin\varphi_0-h\cos\varphi_0)}{1+\sqrt{1-e^2}}\right].
1111: \end{eqnarray}
1112: Putting all terms together and replacing the appropriate terms by
1113: $e_\perp=k\sin\varphi_0-h\cos\varphi_0$,
1114: $e_\parallel=k\cos\varphi_0+h\sin\varphi_0$ and $\ell=1-\sqrt{1-e^2}$,
1115: we get \eqref{lambdaattransit}.
1116: The partial derivatives of \eqref{lambdaattransit} become
1117: \begin{eqnarray}
1118: \frac{\partial\lambda}{\partial k} & = & -\frac{h}{2-\ell}-(1-\ell)\frac{h+(2+e_
1119: \parallel)\sin\varphi_0}{(1+e_\parallel)^2}, \label{e:lambda_k}\\
1120: \frac{\partial\lambda}{\partial h} & = & +\frac{k}{2-\ell}+(1-\ell)\frac{k+(2+e_
1121: \parallel)\cos\varphi_0}{(1+e_\parallel)^2}. \label{e:lambda_h}
1122: \end{eqnarray}
1123:
1124: \section{Tangential velocity and position at the transit}
1125: \label{appendixtangentvelocity}
1126:
1127: It is known from the theory of the two-body problem that the
1128: angular momentum of a body orbiting around a mass of $GM=\mu$
1129: and having an orbit with the semimajor axis of $a$ and eccentricity $e$
1130: is $C=\sqrt{\mu a(1-e^2)}$. Since $C=rv_{\rm tan}$ for all points,
1131: the tangential velocity would be
1132: \begin{equation}
1133: v_{\rm tan}=\frac{C}{r}=\sqrt{\mu a(1-e^2)}\frac{1+e\cos(\varphi_0-\omega)}{a(1-e^2)}\label{vtran1}
1134: \end{equation}
1135: Using Kepler's Third Law, i.e. $\mu=n^2a^3$, the above equation can be
1136: reordered to
1137: \begin{equation}
1138: v_{\rm tan}=an\frac{1+e\cos(\varphi_0-\omega)}{\sqrt{1-e^2}}.\label{vtran2}
1139: \end{equation}
1140: For $\varphi_0=\pi/2$, \eqref{vtran2} becomes
1141: \begin{equation}
1142: v_{\rm tan}=an\frac{1+e\sin\omega}{\sqrt{1-e^2}}.
1143: \end{equation}
1144:
1145: \label{lastpage}
1146:
1147: \end{document}
1148:
1149: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1150: