1: \documentclass[12pt,preprint]{aastex}
2: %\usepackage{lscape,amsmath}
3: %\usepackage[margin=1in]{geometry}
4: \begin{document}
5: \title{Characterizing Long-Period Transiting Planets Observed by {\it Kepler}}
6: \author{Jennifer C.\ Yee and B.\ Scott Gaudi}
7: \affil{Department of Astronomy, The Ohio State University, 140 W. 18th Ave., Columbus, OH 43120}
8:
9: \email{jyee,gaudi@astronomy.ohio-state.edu}
10:
11: \begin{abstract}
12: {\it Kepler} will monitor a sufficient number of stars that it is likely to
13: detect single transits of planets with periods longer than the mission
14: lifetime. We show that by combining the exquisite {\it Kepler} photometry of
15: such transits with precise radial velocity observations taken over a
16: reasonable timescale ($\sim 6$ months) after the transits,
17: and assuming circular orbits, it is
18: possible to estimate the periods of these transiting planets to
19: better than 20\%, for planets with radii greater than that of Neptune,
20: and the masses to within a factor of 2, for planets with masses larger
21: than or about equal to the mass of Jupiter. Using a Fisher matrix analysis, we
22: derive analytic estimates for the uncertainties in the velocity of the
23: planet and the acceleration of the star at the time of transit, which we
24: then use to derive the uncertainties for the planet mass, radius,
25: period, semimajor axis, and orbital inclination. Finally, we explore
26: the impact of orbital eccentricity on the estimates of these
27: quantities.
28: \end{abstract}
29:
30: \keywords{methods: analytical, planetary systems, planets and satellites: general}
31:
32: \section{Introduction}
33: Planets that transit their stars offer us the opportunity to study the
34: physics of planetary atmospheres and interiors, which may help
35: constrain theories of planet formation. From the photometric light
36: curve, we can measure the planetary radius and also the orbital
37: inclination, which when combined with radial velocity (RV) observations,
38: allows us to measure the mass and density of the planet. Infrared
39: observations of planets during secondary eclipse can be used to
40: measure the planet's thermal spectrum \citep{Charbonneau05,Deming05},
41: and spectroscopic observations around the time of transit can constrain the
42: composition of the planet's atmosphere through transmission
43: spectroscopy \citep{Charbonneau02}. Additionally, optical
44: observations of the secondary eclipse can probe a planet's albedo
45: \citep{Rowe07}. All of the known transiting planets are Hot Jupiters or Hot Neptunes,
46: which orbit so close to their parent stars that the stellar flux plays
47: a major role in heating these planets. In contrast, the detection of
48: transiting planets with longer periods ($P > 1\, {\rm yr}$) and
49: consequently lower equilibrium temperatures, would allow us to probe a
50: completely different regime of stellar insolation, one more like that
51: of Jupiter and Saturn whose energy budgets are dominated by their
52: internal heat.
53:
54: Currently, it is not feasible to find long-period, transiting planets
55: from the ground. While RV surveys have found $\sim 110$ planets with
56: periods $\ga 1\, {\rm yr}$, it is unlikely that any of these transit
57: their parent star, given that for a solar-type host, the transit
58: probability is very small, $\wp_{\rm tr}\simeq 0.5\% (P/{\rm yr})^{-2/3}$. Thus the
59: sample of long-period planets detected by RV will need to at least
60: double before a transiting planet is expected. Given the long lead
61: time necessary to detect long-period planets with RV observations, this
62: is unlikely to happen in the near future. Furthermore, the expected
63: transit times for long-period systems are relatively uncertain, which
64: makes the coordination and execution of photometric follow-up
65: observations difficult. Putting all this together, radial velocity
66: surveys are clearly an inefficient means to search for long-period,
67: transiting planets.
68:
69: Ground-based photometric transit surveys are similarly problematic. A
70: long-period system must be monitored for a long time because
71: typically, at least two transits must be observed in order to measure
72: the period of the system. Furthermore, if there are a couple of days
73: of bad weather at the wrong time, the transit event will be missed
74: entirely, rendering years of data worthless. Thus transit surveys for
75: long-period planets require many years of nearly uninterrupted
76: observations. Additionally, many thousands of suitable stars must be
77: monitored to find a single favorably-inclined system.
78:
79: Because of its long mission lifetime ($L = 3.5\, {\rm yrs}$),
80: nearly continuous observations, and large number of target stars ($N \simeq
81: 10^5$), the {\it Kepler} satellite \citep{Borucki04, Basri05} has a unique opportunity to
82: discover long-period transiting systems. Not only will {\it Kepler}
83: observe multiple transits of planets with periods up to the mission
84: lifetime, but it is also likely to observe single transits of planets
85: with periods longer than the mission lifetime. As the period of a
86: system increases beyond $L/2$, the probability of observing more than
87: one transit decreases until, for periods longer $L$, only one transit
88: will ever be observed. For periods longer than $L$, the probability
89: of seeing a single transit diminishes as $P^{-5/3}$. Even with only a
90: single transit observation from {\it Kepler}, these long-period
91: planets are invaluable.
92:
93: We show that planets with periods longer than the mission lifetime
94: will likely be detected and can be characterized using the {\it
95: Kepler} photometry and precise radial velocity
96: observations. Furthermore, this technique can be applied to planets
97: that will transit more than once during the {\it Kepler} mission, so
98: that targeted increases in the time-sampling rate can be made at the
99: times of subsequent transits.
100:
101: In $\S$\ref{sec:np} we calculate the number of one- and
102: two-transit systems {\it Kepler} is expected to observe. We give a
103: general overview of how the {\it Kepler} photometry and radial
104: velocity follow-up observations can be combined to characterize a
105: planet in $\S$\ref{sec:basic}. We discuss in $\S$\ref{sec:lc} the expected
106: uncertainties in the light curve observables, and
107: we relate these to uncertainties in physical quantities, such as the
108: period, that can be derived from the transit light
109: curve. $\S$\ref{sec:rv} describes the expected uncertainties
110: associated with the RV curve and how they influence the uncertainty in
111: the mass of the planet. $\S$\ref{sec:e} discusses the potential impact
112: of eccentricity on the ability to characterize long-period planets. We
113: summarize our conclusions in $\S$\ref{sec:conclude}. Details of the
114: derivations are reserved for an appendix ($\S$\ref{sec:append}).
115:
116: \section{Expected Number of Planets}
117: \label{sec:np}
118:
119: The number of single-transit events {\it Kepler} can expect to find is
120: the integral of the probability that a given system will be favorably
121: inclined to produce a transit, multiplied by the probability that a
122: transit will occur during the mission, convolved with a distribution
123: of semimajor axes, and normalized by the expected frequency of
124: planets and the number of stars that {\it Kepler} will observe,
125:
126: \begin{equation}
127: \label{eqn:np}
128: N_{\rm tot} = N_{\star}\int\wp_{\rm tr}\wp_Lf(a) da .
129: \end{equation}
130: Here $N_{\rm tot}$ is the total number of transiting systems expected,
131: $N_{\star}$ is the number of stars being monitored, $\wp_{\rm tr}$ is the
132: probability that the system is favorably inclined to produce a transit
133: (the transit probability), $\wp_L$ is the probability that a
134: transit will occur during the mission, and $f(a)$ is some distribution
135: of the planet semimajor axis $a$, normalized to the expected
136: frequency of planets.
137:
138: Assuming a circular orbit, the transit probability is simply,
139: \begin{equation}
140: \wp_{\rm tr} = \frac{R_{\star}}{a}= \left(\frac{4 \pi^2}{G}\right)^{1/3}M_{\star}^{-1/3} R_{\star} P^{-2/3},
141: \end{equation}
142: where $R_{\star}$ and $M_{\star}$ are the radius and mass of the parent star, and $P$ is the period of the system.
143:
144: We consider both the probability of observing exactly one
145: transit, $\wp_{L,1}$, and the probability of observing exactly two
146: transits, $\wp_{L,2}$. These probabilities are
147: \begin{equation}
148: \label{eqn:1trans}
149: \wp_{L,1} =\left\{
150: \begin{array}{ll}
151: \frac{2P}{L} - 1 & \frac{L}{2} \le P \le L , \\
152: \frac{L}{P} & P \ge L .
153: \end{array}\right.
154: \end{equation}
155: \begin{equation}
156: \label{eqn:2trans}
157: \wp_{L,2} = \left\{
158: \begin{array}{ll}
159: \frac{4P}{L} - 1 & \frac{L}{4} \le P \le \frac{L}{2}, \\
160: 2 -\frac{2P}{L} & \frac{L}{2} \le P \le L .
161: \end{array}\right.
162: \end{equation}
163: Thus there will be a range in periods from $L/2$ to
164: $L$ where it is possible to get either one or two transits during the
165: mission. If we assume a mission lifetime\footnote{Here and throughout this paper, we use the characteristics of the {\it Kepler} mission provided by the website, http://kepler.nasa.gov/.}
166: of $L = 3.5$ years and a solar-type
167: star ($R_{\star}=R_{\odot}, M_{\star}=M_{\odot}$), we can find the
168: total probability ($\wp_{\rm tr}\wp_L$) of observing a planet that
169: transits {\it and} exhibits exactly one or exactly two transits as
170: a function of period or semimajor axis. These probability distributions
171: as functions of $a$ are shown in the middle panel of Fig. \ref{fig:np}.
172:
173: For $P \geq L$, the total probability, $\wp_{\rm tot}$, of observing a single transit is
174: \begin{equation}
175: \wp_{\rm tot} \equiv \wp_{\rm tr}\wp_L = 0.002\left(\frac{R_{\star}}{R_{\odot}}\right)\left(\frac{M_{\star}}{M_{\odot}}\right)^{-1/3}\left(\frac{P}{3.5 {\rm\, yrs}}\right)^{-5/3}\left(\frac{L}{3.5 {\rm\, yrs}}\right) .
176: \end{equation}
177: Thus the probability that
178: the planet will transit {\it and} that a
179: single transit will occur during the mission is generally
180: small for periods at the mission lifetime, and beyond this is a strongly
181: decreasing function of the period, $\propto P^{-5/3}$
182: or $\propto a^{-5/2}$.
183:
184: The observed distribution of semimajor axes for planets with minimum
185: masses $m_{\rm p}\sin i \ge 0.3~M_{\rm Jup}$ is shown in the top panel of
186: Figure \ref{fig:np}, where $m_{\rm p}$ is the mass of the planet, $i$ is the
187: orbital inclination, and the data are taken from the ``Catalog of
188: Nearby Exoplanets'' of \citet{Butler06}\footnote{The catalog is
189: available from http://exoplanets.org/planets.shtml. It was accessed 2008 May 9.}. Note that we
190: have included multiple-planet systems. We
191: normalized the distribution by assuming that 8.5\% of all stars have
192: planets with $m_{\rm p}\sin i \ge 0.3~M_{\rm Jup}$ and $a\le 3~{\rm AU}$
193: \citep{Cumming08}. We can extrapolate this distribution to larger semimajor
194: axes by observing that it is approximately constant as a function of
195: $\log a$ and adopting the value found by \citet{Cumming08}, $dN/d\log{a} = 4.3\%$.
196: Using this extrapolation,
197: we predict $13.4\%$ and $16.4\%$ of stars have
198: a $m_{\rm p}\sin i \ge 0.3~M_{\rm Jup}$ planet within $10~{\rm AU}$ and
199: $20~{\rm AU}$, respectively, in agreement with \citet{Cumming08}.
200:
201: We find the number of planets expected for {\it Kepler} by convolving this normalized
202: distribution with the probability distribution and multiplying by $N_{\star} = 100,000$. We convolve the distribution of semimajor axes with
203: the total transit probability distribution by weighting the probability of observing
204: a transit for each member of the bin to determine an overall
205: probability for the bin. The bottom panel of Fig. \ref{fig:np} shows
206: the resulting distribution of one- and two-transit
207: systems. Integrating this distribution over all semimajor axes out to the limit of current RV observations predicts that {\it Kepler} will observe a total of 4.0 single-transit
208: systems and 5.6 two-transit systems during the mission lifetime of $L = 3.5 \, {\rm yrs}$. If we include the extrapolation
209: of the observed distribution to larger semimajor axes, the integrated number of single-transit systems
210: increases to 5.7; the effect is modest because the probability of
211: observing a transit declines rapidly with increasing semimajor axis
212: ($\wp_{\rm tot} \propto a^{-5/2}$).
213:
214: Thus, {\it Kepler} is likely to detect at least handful of single-transit events.
215: It is important to bear in mind that this is a lower
216: limit. First, our estimate does not include planets with $m_{\rm p}\sin i \la 0.3~M_{\rm Jup}$,
217: primarily because RV surveys are substantially incomplete for low-mass,
218: long-period planets. As we will show, {\it Kepler} will be able to
219: constrain the periods of planets with radii as small as that of Neptune
220: with the detection of a single transit. Indeed, microlensing surveys indicate
221: that cool, Neptune-mass planets are common \citep{Beaulieu06,Gould06}.
222: Second, \citet{Barnes07} and \citet{Burke08} showed that a distribution in eccentricity
223: increases the probability that a planet will transit its parent star.
224: Finally, if the {\it Kepler} survey lifetime is extended, the number
225: of detections of long-period transiting planets will also increase.
226:
227: One might assume that a single-transit event must simply be discarded from the sample
228: because the event is not confirmed by a second transit, and the period
229: cannot be constrained by the time between successive transits.
230: Although we do not expect to see many of them, long-period
231: transiting planets are precious, as the information we could potentially
232: gain by observing these systems when they transit in the future would allow us
233: to greatly enhance our understanding of the physical properties of outer giant planets, and
234: in particular would allow us to compare them directly to our own solar system giants. Thus, these
235: single transit events should be saved, if at all possible.
236:
237: \begin{figure}
238: \begin{center}
239: \includegraphics{f1.eps}
240: \end{center}
241:
242: \caption{({\it top})
243: The observed distribution of semimajor axes for planets with minimum
244: masses $m_{\rm p}\sin i \ge 0.3~M_{\rm Jup}$.
245: The shaded region indicates $a \ge 10^{0.5}\,\mathrm{AU} \approx 3\,{\rm AU}$, where we expect
246: the sample to suffer from incompleteness. The dashed line indicates our extrapolation of the
247: distribution, found by assuming the constant
248: distribution in $\log a$ from \citet{Cumming08}. ({\it middle})
249: The total transit probability. The solid line shows the probability that a planet
250: transits {\it and} exhibits a single transit during the {\it Kepler} mission
251: lifetime. The dotted line shows the same probability for two
252: transits. ({\it bottom}) The expected number of one and two-transit events. The
253: probability distribution has been convolved with the observed fraction
254: of planets and multiplied by the number of stars Kepler will observe (100,000) to give the number of single transit systems (solid line)
255: and two transit systems (dotted line) that can be expected. The
256: dashed line shows the convolution of the single transit probability
257: distribution with the extrapolation to larger semimajor
258: axes. \label{fig:np}}
259: \end{figure}
260:
261: \section{Characterizing a Planet}
262: \label{sec:basic}
263:
264: With a few basic assumptions, {\it Kepler} photometry of a transiting planet
265: can provide a period for the system that can be combined with precise
266: radial velocity measurements to estimate a mass for the planet. Here we
267: briefly sketch the basics of how these properties can be estimated and what their uncertainties are, and we provide details in the following sections.
268: We assume the planet is much less massive than the star, $m_{\rm p} \ll M_{\star}$, and circular orbits. We consider the impact of eccentricity
269: on our conclusions in \S\ref{sec:e}.
270:
271: Assuming no limb-darkening and assuming the out-of-transit flux
272: is known perfectly\footnote{As we will show,
273: including limb-darkening and the uncertainty in the out-of-transit flux does not change the following discussion qualitatively.},
274: a transit can be characterized by four observables,
275: namely the fractional depth of the transit $\delta$, the full-width half-maximum of the
276: transit $T$, the ingress/egress duration $\tau$, and the time of the center of transit $t_c$.
277: These can be combined to estimate the instantaneous velocity of the
278: planet at the time of transit,
279: \begin{equation}
280: v_{\rm tr,p} = 2R_{\star}\left(\frac{\sqrt{\delta}}{T\tau}\right)^{1/2}.
281: \label{eqn:vtr}
282: \end{equation}
283: For a circular orbit, $v_{\rm tr,p}$ can be used to estimate
284: the period of the planet
285: \begin{equation}
286: P = \frac{2\pi a}{v_{\rm tr,p}} = \frac{G\pi^2}{3}\rho_{\star}\left(\frac{T\tau}{\sqrt{\delta}}\right)^{3/2},
287: \end{equation}
288: where we have employed Kepler's third law assuming that the
289: planet's mass is much smaller than the star's mass. Thus, if the stellar
290: density $\rho_{\star}$ is known, then the planet's period
291: can be estimated from photometry of a single transit \citep{Seager03}. Note that there is a degeneracy between $\rho_{\star}$ and $P$, so with a single transit $\rho_{\star}$ must be determined by other means in order to derive an estimate of $P$. $\rho_{\star}$ can be estimated via spectroscopy combined with theoretical isochrones, or asteroseismology.
292: The total uncertainty in the period is the quadrature sum of the contribution
293: from the estimate of $\rho_{\star}$ and the contribution from the {\it Kepler} photometry
294: (i.e., the uncertainty at fixed $\rho_{\star}$),
295: \begin{equation}
296: \left(\frac{\sigma_P}{P}\right)_{\rm tot}^2 =
297: \left(\frac{\sigma_{\rho_{\star}}}{\rho_{\star}}\right)^2 +
298: \left(\frac{\sigma_P}{P}\right)_{\rm Kep}^2.
299: \end{equation}
300: The contribution from {\it Kepler} is dominated by the uncertainty in the ingress/egress time $\tau$,
301: \begin{equation}
302: \left(\frac{\sigma_P}{P}\right)_{\rm Kep}^2
303: \approx \frac{9}{4} \left(\frac{\sigma_{\tau}}{\tau}\right)^2
304: \approx Q^{-2}\left(\frac{27T}{2\tau}\right),
305: \end{equation}
306: where we have assumed $\tau \ll T$, and that the number of points taken out of
307: transit is much larger than the number taken during the transit (and thus the out-of-transit
308: flux is known essentially perfectly).
309: Here $Q$ is approximately equal to the total signal-to-noise ratio of the transit,
310: \begin{equation}
311: Q \equiv (\Gamma_{\rm ph} T)^{1/2} \delta,
312: \end{equation}
313: where $\Gamma_{\rm ph}$ is the photon collection rate. For {\it Kepler} we assume,
314: \begin{equation}
315: \Gamma_{\rm ph}= 7.8 \times 10^8~{\rm hr^{-1}}\, 10^{-0.4(V-12)},
316: \end{equation}
317: and thus {\it Kepler}
318: will detect a single transit
319: with a signal-to-noise ratio of
320: \begin{equation}
321: Q \simeq 1300\,
322: \left(\frac{R_{\star}}{R_{\odot}}\right)^{-3/2}
323: \left(\frac{M_{\star}}{M_{\odot}}\right)^{-1/6}
324: \left(\frac{r_{\rm p}}{R_{\rm Jup}}\right)^2
325: \left(\frac{P}{3.5 {\rm\, yrs}}\right)^{1/6}
326: 10^{-0.2(V-12)},
327: \end{equation}
328: where $r_{\rm p}$ is the radius of the planet. For a Neptune-size planet with these fiducial parameters,
329: $Q \simeq 150$.
330:
331: {\it Kepler's} contribution to the uncertainty in the period is,
332: \begin{equation}
333: \left(\frac{\sigma_P}{P}\right)_{\rm Kep}^2=
334: 7.7\times10^{-5}10^{0.4(m_v-12)}
335: \left(\frac{R_{\star}}{R_{\odot}}\right)^4
336: \left(\frac{M_{\star}}{M_{\odot}}\right)^{1/3}
337: \left(\frac{R_{\rm Jup}}{r_{\rm p}}\right)^5
338: \left(\frac{3.5 {\rm\, yrs}}{P}\right)^{1/3},
339: \end{equation}
340: where we have assumed the impact
341: parameter $b=0$.
342: Due to the strong scaling with $r_{\rm p}$ and relatively
343: weak scaling with $P$, there are essentially two regimes. For $r_{\rm p} \ga R_{\rm Nep}$,
344: the uncertainty in $P$ is dominated by uncertainties in the estimate of $\rho_{\star}$,
345: which is expected to be ${\cal O} (10\%)$, whereas for $r_{\rm p} < R_{\rm Nep}$, it is very
346: difficult to estimate the period due to the uncertainty in $\tau$ from the {\it Kepler} photometry.
347:
348: Once the period is known from the photometry, the mass of the planet,
349: $m_{\rm p}$, can be measured with radial velocity observations. For a
350: circular orbit, and for observations spread out over a time that is short
351: compared to $P$, the radial velocity of the star can be expanded about the time
352: of transit, and so approximated by the velocity at the time of transit $v_0$, plus
353: a constant acceleration $A_{\star}$,
354: \begin{equation}
355: v_{\star} \approx v_0 - A_{\star}(t-t_c),
356: \end{equation}
357: where $A_{\star}=2\pi K_{\star}/P$, and
358: \begin{equation}
359: K_{\star} = \left(\frac{2 \pi G}{PM_{\star}^2}\right)^{1/3} m_{\rm p} \sin i
360: \end{equation}
361: is the stellar radial velocity semi-amplitude. Thus,
362: \begin{equation}
363: m_{\rm p} = A_{\star}\left(\frac{M_{\star}^2}{G}\right)^{1/3}\left(\frac{P}{2\pi}\right)^{4/3}
364: = \frac{1}{16G}g_{\star}^2A_{\star}\left(\frac{T\tau}{\sqrt{\delta}}\right)^2,
365: \end{equation}
366: where $g_{\star}$ is the surface gravity of the star, and we have assumed $\sin\, i = 1$.
367:
368: The uncertainty in the planet mass has contributions from three distinct sources:
369: the uncertainty in $g_{\star}$, which can be estimated from spectroscopy, the uncertainty in $A_{\star}$,
370: which is derived from RV observations after the transit, and the uncertainties in
371: $T,\, \tau,$ and $\delta$, which are derived from {\it Kepler} photometry. In fact, the uncertainty
372: in $\tau$ dominates over the uncertainties in $T$ and $\delta$. Therefore, we may write,
373: \begin{equation}
374: \label{eqn:sigB}
375: \left(\frac{\sigma_{m_{\rm p}}}{m_{\rm p}}\right)^2 = 4\left(\frac{\sigma_{g_{\star}}}{g_{\star}}\right)^2+
376: \left(\frac{\sigma_{A_{\star}}}{A_{\star}}\right)^2+
377: 4\left(\frac{\sigma_{\tau}}{\tau}\right)^2.
378: \end{equation}
379: For reasonable assumptions and long-period planets ($P \ga 1~{\rm yr}$), we find that the uncertainty in $A_{\star}$ dominates
380: over the uncertainty in $\tau$.
381:
382: Assuming that $N$ equally-spaced RV measurements with
383: precision $\sigma_{\rm RV}$ are taken over a time period $T_{\rm tot}$ after the transit, the uncertainty in $A_{\star}$ is,
384: \begin{eqnarray}
385: \left(\frac{\sigma_{A_{\star}}}{A_{\star}}\right)^2 &\simeq& \frac{12 \sigma_{\rm RV}^2}{A_{\star}^2 T_{\rm tot}^2 N},\nonumber\\
386: &\simeq& 0.85\left(\frac{\sigma_{\rm RV}}{10 {\rm\, m\, s}^{-1}}\right)^2\left(\frac{3 {\rm\, mos}}{T_{\rm tot}}\right)^2\left(\frac{20}{N}\right)\left(\frac{M_{\rm Jup}}{m_{\rm p}}\right)^2\left(\frac{M_{\star}}{M_{\odot}}\right)^{4/3}\left(\frac{P}{3.5 {\rm\, yrs}}\right)^{8/3},
387: \label{eqn:sigBscale}
388: \end{eqnarray}
389: where we have assumed $N\gg 1$. Thus, radial
390: velocity observations combined with {\it Kepler} photometry can confirm the
391: planetary nature of a Jupiter-sized planet in a relatively short time
392: span. An accurate ($\la 10\%$) measurement of the mass for a Jupiter-mass planet, or even a rough
393: characterization of the mass for a Neptune-mass planet, will require either more measurements,
394: or measurements with substantially higher RV precision.
395: Of course, additional radial velocity observations over a time span comparable to $P$
396: will further constrain the mass and period.
397:
398: In $\S$\ref{sec:lc} and $\S$\ref{sec:rv}, we derive the
399: above expressions for the uncertainty in the mass and the period of the
400: planet using a Fisher information analysis.
401:
402: \section{Estimating the Uncertainty in P}
403: \label{sec:lc}
404:
405: \subsection{Uncertainties in the Light Curve Observables}
406: \label{sec:models}
407: In the absence of significant limb-darkening, a transit light curve can be approximated by
408: a trapezoid that is described by the five parameters $t_c$, $T$, $\tau$,
409: $\delta$, and $F_0$, where
410: $F_0$ is the out-of-transit flux. Figure \ref{fig:lc} shows this simple trapezoidal model and labels the relevant
411: parameters. Mathematically, the flux $F$ as a function of time $t$ is given by,
412: \begin{equation}
413: \label{eqn:lc}
414: F(t)=\left\{
415: \begin{array}{ll}
416: F_0, & t < t_1 ,\\
417: F_0-\delta(t -[t_c-T/2-\tau/2])/\tau, & t_1 \le t \le t_2 ,\\
418: F_0-\delta, & t_2 < t < t_3 ,\\
419: F_0-\delta(1-[(t -[t_c-T/2-\tau/2])/\tau]), & t_3 \le t \le t_4 ,\\
420: F_0, & t_4 < t ,
421: \end{array}\right.
422: \end{equation}
423: where $t_1$ -- $t_4$ are the points of contact. We also
424: define $D$ to be the total duration of the observations.
425: This model is fully differentiable
426: and can be used with the Fisher matrix formalism to derive exact
427: expressions for the uncertainties in the parameters $t_c$, $T$, $\tau$,
428: $\delta$, and $F_0$. We note that the definition for $\delta$
429: we use here differs slightly from the definition we adopted in \S\ref{sec:basic}.
430: Here $\delta$ is the depth of the transit (e.g., in units of flux), rather than the fractional depth.
431: These two definitions differ by a factor of $F_0$ such that
432: $\delta_{frac}=\delta_{flux}/F_0$. However, when $D \gg T$ (as will be the case for {\it Kepler}), the uncertainty in $F_0$
433: is negligible, and
434: so one may define $F_0=1$, thus making these two parametrizations equivalent.
435:
436:
437: Figure \ref{fig:models} compares this simple model with the exact
438: model computed using the formalism of \citet{Mandel02}, in this case
439: for a transit of the Sun by Neptune with a 3.5-year period and an
440: impact parameter of $b=0.2$. We see that the trapezoidal model
441: provides an excellent approximation to the exact light curve, and as
442: we will show the analytic parameter uncertainties are quite accurate. We
443: also show the case of significant limb-darkening as expected
444: for a Sun-like star observed in the {\it R}-band (similar to the {\it Kepler} bandpass). In this case, the
445: match is considerably poorer, but nevertheless we will show the analytic parameter
446: uncertainties we derive assuming the simple trapezoidal model still
447: provide useful estimates (see \citealt{Carter08} for a more thorough
448: discussion).
449:
450: We derive exact expressions for the uncertainties in $t_c$, $T$, $\tau$,
451: $\delta$, and $F_0$ by applying the Fisher matrix formalism to the
452: simple light curve model (a detailed explanation of the Fisher matrix
453: is given by \citealt{Gould03}). The full details of the derivation are
454: given in the appendix ($\S$\ref{sec:append}). In the case of the
455: long-period transiting planets that will be observed by {\it Kepler}, we can
456: simplify the full expression by making a couple of assumptions. We
457: assume that the flux out of transit, $F_0$, is known to infinite
458: precision and that $\tau \ll T \ll D$. The uncertainties in the remaining
459: parameters are,
460: \begin{eqnarray}
461: \label{tc} & \sigma_{t_c} & = Q^{-1}\sqrt{\frac{T\tau}{2}} ,\\
462: & \frac{\sigma_T}{T} & = Q^{-1}\sqrt{\frac{2\tau}{T}} ,\\
463: & \frac{\sigma_{\tau}}{\tau} & = Q^{-1}\sqrt{\frac{6T}{\tau}} ,\\
464: \label{d} & \frac{\sigma_{\delta}}{\delta} & = Q^{-1}.
465: \end{eqnarray}
466:
467: We tested the accuracy of these expressions
468: using Monte Carlo simulations. We generate 1000
469: light curves for a given set of orbital parameters. We assume that the
470: planet orbits a solar-type star in a circular orbit with an impact
471: parameter of 0.2. We repeat the analysis for four of the solar-system
472: planets: Jupiter, Saturn, Neptune, and Earth. We assume the expected
473: photometric precision for {\it Kepler} for a stellar apparent
474: magnitude of $V=12$ (a representative magnitude for {\it Kepler}'s stellar sample), a sampling
475: rate of one per $30\, \mathrm{min}$, and a total duration for
476: the observations of $D = 200$ hours. Then we fit for the parameters
477: $t_c$, $T$, $\tau$, $\delta$, and $F_0$ using a down-hill simplex
478: method. The uncertainty in each parameter is taken to be the standard
479: deviation in the distribution of the fits for that parameter. A
480: comparison of the analytic expressions and the Monte Carlo simulations
481: is shown in Fig. \ref{fig:sigobs}, where we have plotted the
482: uncertainties as a function of period for a planet with radius $1
483: R_{\rm Nep}$ crossing a star with radius $1 R_{\odot}$ at an impact
484: parameter $b=0.2$. We find that the Fisher matrix approximation breaks
485: down as the number of points during ingress/egress becomes small.
486:
487: We also considered expected uncertainties for the exact
488: uniform source and limb-darkened models for
489: the transit light curve. \citet{Mandel02} give the full solution for
490: a transit involving two spherical bodies and provide code for
491: calculating the transit light curve. They also include a
492: limb-darkened solution. These models may be parametrized by the same
493: observables described in our simplified model. The full and
494: limb-darkened light curves are plotted in Fig. \ref{fig:models} as
495: the dotted and dashed curves, respectively. We use the
496: limb-darkening coefficients from \citet{Claret00} closest to
497: observations of the Sun ($T_{\rm eff} = 5750~{\rm K}$, $\log(g_{\star})= 4.5~{\rm\, cm\, s}^{-1}$,
498: $[M/H] = 0.0$) in the R-band. We apply the Fisher matrix
499: method to numerical derivatives of these model light curves to compare
500: the uncertainties for the full and limb-darkened solutions with the
501: analytic uncertainties for the trapezoidal model. A
502: comparison of the uncertainties from the different models is shown
503: in Fig. \ref{fig:sigobs}. We find that the simplified model is a good
504: approximation for the exact, uniform-source transit model. While the comparison
505: is less favorable with the limb-darkening model, the uncertainties do not differ by more
506: than a factor of a few.
507:
508: We also compared how the uncertainties in the transit observables
509: varied with impact parameter $b$ for the three models, because varying $b$ will affect the relative sizes of $T$ and $\tau$. These comparisons
510: are shown in Fig. \ref{fig:impact} for Neptune in a circular orbit
511: around the Sun with a 3.5-year period. As can be seen from the figure,
512: variations in $b$ have little impact in the uncertainties in the
513: observables for $b \lesssim 0.8$.
514:
515: \begin{figure}
516: \includegraphics[width=6in]{f2.eps}
517: \caption{
518: The simplified model of a transit light curve showing the points of
519: contact ($t_1, t_2, t_3, t_4$), the transit depth ($\delta$),
520: the ingress/egress time ($\tau$), the FWHM
521: duration of the transit ($T$), the total duration of observations
522: ($D$), and the time of the transit center ($t_c$). \label{fig:lc}}
523: \end{figure}
524:
525: \begin{figure}
526: \includegraphics[width=6in]{f3.eps}
527: \caption{
528: Three different models of a transit light curve. The solid line shows
529: the simplified trapezoidal model, the dotted line (barely visible) shows the exact,
530: uniform-source light curve, and the
531: dashed line includes the effects of limb-darkening. The exact uniform-source and
532: limb-darkened light curves are calculated using \citet{Mandel02}. These light curves were generated for a Neptune
533: analog orbiting a solar analog at a period of $P=3.5~{\rm yrs}$ and an impact parameter $b=0.2$. Typical error bars for Kepler assuming $V=12.0$ and 30-min sampling (solid) or 20-min sampling (dotted) are indicated.
534: \label{fig:models}}
535: \end{figure}
536:
537: \begin{figure}
538: \includegraphics[width=5.5in]{f4.eps}
539: \caption{
540: Uncertainty in observables as a function of period. The uncertainties
541: in $t_c$, $T$, $\tau$, and $\delta$ are plotted versus period for Neptune orbiting the Sun with an impact parameter of 0.2. The
542: different line styles indicate the trapezoidal transit model (solid
543: line), the full solution (dotted line), and a model including the
544: effects of limb-darkening (dashed line). Monte-Carlo simulations
545: of the trapezoidal model using 30-minute sampling are shown as crosses. The gray crosses show simulations for 20-minute sampling. The dotted line in the
546: bottom panel is not visible because it is nearly identical to the
547: solid line. The vertical dotted lines mark where the ingress/egress duration $\tau$ is equal to the sampling rate $dt$, as well as where $\tau=2dt$ for 30-minute sampling.\label{fig:sigobs}}
548: \end{figure}
549:
550: \begin{figure}
551: \includegraphics[width=5.5in]{f5.eps}
552: \caption{
553: Uncertainty in observables as a function of impact parameter. The
554: uncertainties in $t_c$, $T$, $\tau$, and $\delta$ are plotted versus impact
555: parameter for Neptune orbiting the Sun with a period of 3.5 years. The
556: line styles are the same as for
557: Fig. \ref{fig:sigobs}. As in Fig. \ref{fig:sigobs}, the dotted line in the bottom panel is not visible because it is nearly identical to the solid line.
558: \label{fig:impact}}
559: \end{figure}
560:
561: \subsection{Uncertainties in Quantities Derived From the Light Curve}
562:
563: Assuming the planet is in a circular orbit, and assuming
564: the stellar density, $\rho_{\star}$, is
565: estimated from independent information, \citet{Seager03} demonstrated
566: that the planet period, $P$, and impact parameter, $b$, can derived from a single observed transit,
567: with sufficiently precise photometry. In fact, provided
568: independent estimates of $M_{\star}$ and $R_{\star}$ are available,
569: then it is also possible to derive the planet radius, $r_{p}$, semimajor
570: axis, $a$, and instantaneous velocity at the time of transit,
571: $v_{\rm tr,p}$.
572:
573: The uncertainties in these physical parameters can be
574: calculated through standard error propagation. Equations for these
575: quantities and their uncertainties are given in
576: Eqs. \ref{eqn:rp}--\ref{eqn:p} (Note that we calculate $b^2$ instead
577: of $b$ because numerical methods have difficulty when
578: $T\sqrt{\delta}/\tau \approx 1$). In each case below,
579: the second approximate uncertainty equation is calculated by substituting the variances and
580: covariances of the observables $\tau,\,T$, and $\delta$, and simplifying under the
581: assumption that $\tau \ll T \ll D$.
582:
583: \begin{eqnarray}
584: \label{eqn:rp}
585: r_{\rm p} &= &R_{\star}\sqrt{\delta}\\
586: \sigma_{r_{\rm p}}^2 &= &r_{\rm p}^2\left(\frac{1}{R_{\star}}\sigma_{R_{\star}}^2+\frac{1}{4\delta^2}\sigma_{\delta}^2\right)\nonumber\\
587: & \simeq &r_{\rm p}^2\left(\frac{1}{R_{\star}}\sigma_{R_{\star}}^2+\frac{1}{4}Q^{-2}\right)\nonumber\\
588: && \nonumber\\
589: b^2 &=& 1-\frac{T\sqrt{\delta}}{\tau}\\
590: \sigma_{b^2}^2 &=& \left[\frac{\delta}{\tau^2}\sigma_T^2 + \frac{T^2\delta}{\tau^4}\sigma_{\tau}^2 + \frac{T^2}{4\tau^2\delta}\sigma_{\delta}^2 -\frac{2T\delta}{\tau^3}\sigma_{T\tau}^2+\frac{T}{\tau^2}\sigma_{T\delta}^2-\frac{T^2}{\tau^3}\sigma_{\tau\delta}^2 \right]\nonumber\\
591: &\simeq& T\delta Q^{-2}\left(\frac{6T^2}{\tau^3}\right)\nonumber\\
592: && \nonumber\\
593: a &=& \frac{g_{\star}}{4}\left(\frac{T\tau}{\sqrt{\delta}}\right)\\
594: \sigma_a^2 &=& a^2\left[\frac{1}{g_{\star}^2}\sigma_{g_{\star}}^2+ \frac{1}{T^2}\sigma_T^2 + \frac{1}{\tau^2}\sigma_{\tau}^2+ \frac{1}{4\delta^2}\sigma_{\delta}^2+\frac{2}{T\tau}\sigma_{T\tau}^2 -\frac{1}{T\delta}\sigma_{T\delta}^2 - \frac{1}{\tau\delta}\sigma_{\tau\delta}^2 \right]\nonumber\\
595: &\simeq& a^2\left[\frac{1}{g_{\star}^2}\sigma_{g_{\star}}^2+Q^{-2}\left(\frac{6T}{\tau}\right)\right]\nonumber\\
596: && \nonumber\\
597: v_{\rm tr,p} &=& 2R_{\star}\left(\frac{\sqrt{\delta}}{T\tau}\right)^{1/2}\\
598: \sigma_{v_{\rm tr,p}}^2 &=& v_{\rm tr,p}^2\left[\frac{1}{R_{\star}^2}\sigma_{R_{\star}}^2+\frac{1}{4T^2}\sigma_T^2+\frac{1}{4\tau^2}\sigma_{\tau}^2+\frac{1}{16\delta^2}\sigma_{\delta}^2+\frac{1}{2T\tau}\sigma_{T\tau}^2-\frac{1}{4T\delta}\sigma_{T\delta}^2-\frac{1}{4\tau\delta}\sigma_{\tau\delta}^2\right]\nonumber\\
599: &\simeq& v_{\rm tr,p}^2\left[\frac{1}{R_{\star}^2}\sigma_{R_{\star}}^2 + Q^{-2}\left(\frac{3T}{2\tau}\right)\right]\nonumber\\
600: && \nonumber\\
601: \label{eqn:p}
602: P &=& \frac{\pi G}{4}\left(\frac{4}{3}\pi\rho_{\star}\right)\left(\frac{T\tau}{\sqrt{\delta}}\right)^{\frac{3}{2}}\\
603: \sigma_P^2 &=& P^2\left[ \frac{1}{\rho^2_{\star}}\sigma_{\rho_{\star}}^2 + \frac{9}{4T^2}\sigma_T^2 + \frac{9}{4\tau^2}\sigma_{\tau}^2+ \frac{9}{16\delta^2}\sigma_{\delta}^2+\frac{9}{2T\tau}\sigma_{T\tau}^2 -\frac{9}{4T\delta}\sigma_{T\delta}^2 - \frac{9}{4\tau\delta}\sigma_{\tau\delta}^2\right]\nonumber\\
604: &\simeq& P^2\left[\frac{1}{\rho^2_{\star}}\sigma_{\rho_{\star}}^2 +Q^{-2}\left(\frac{27T}{2\tau}\right) \right]\nonumber
605: \end{eqnarray}
606:
607: Of course, the quantities $R_{\star}$, $g_{\star}$, and $\rho_{\star}$
608: and their uncertainties must be estimated from some other external
609: source of information, such as spectroscopy or theoretical isochrones.
610: We can generally expect that the fractional uncertainties on these
611: quantities will be of order $10\%$. For long-period ($P\ga L=
612: 3.5~{\rm yrs}$), Jupiter-sized planets, $Q \gg
613: 1$, and thus the variances and
614: covariances of the transit observables will be small in comparison to
615: the expected uncertainties on $R_{\star}, \rho_{\star}$, and $g_{\star}$.
616:
617: The fractional uncertainties in these derived parameters are plotted in Fig. \ref{fig:deriv}. We compared the analytic uncertainties given in Eqs. \ref{eqn:rp}--\ref{eqn:p} to the Monte Carlo simulations described in Sec. \ref{sec:models}. The fractional uncertainty in each parameter from the Monte Carlo simulations is one-half the range of the middle 68\% of the data divided by the median. Notice that as $\tau$ decreases from $2dt$ to $dt$, the simulations become increasingly disparate from the theoretical expectations. $\sigma_{r_{\rm p}}/r_{\rm p}$ does not show this behavior because it does not depend on $\tau$. In this regime, as $\tau$ decreases, it becomes increasingly probable that only one point will be taken during ingress leaving $\tau$ relatively unconstrained and increasing its uncertainty and the uncertainty of quantities that depend on it.
618:
619: We now consider in detail the uncertainty in the estimated period.
620: Given that $Q \propto T^{1/2} \delta$, we have that
621: $Q^{-2}(T/\tau) \propto (\delta^2 \tau)^{-1}$. Since
622: $\tau \propto r_{\rm p} P^{1/3}$ and $\delta \propto r_{\rm p}^2$, the contribution to the uncertainty in the
623: period due to the {\it Kepler} photometry is a much stronger
624: function of the planet's radius than its period,
625: $Q^{-2}(27T/2\tau) \propto r_{\rm p}^{-5} P^{-1/3}$.
626: As a result, for a fixed stellar radius $R_{\star}$,
627: the ability to accurately estimate $P$ depends almost entirely on $r_{\rm p}$.
628: Furthermore, because the scaling with $r_{\rm p}$ is so strong,
629: there are essentially two distinct regimes: for large $r_{\rm p}$,
630: the uncertainty in $P$ is dominated by the uncertainty
631: in $\rho_{\star}$, whereas for small $r_{\rm p}$, the period cannot be estimated.
632: The boundary between these two regimes is
633: where the contribution to the uncertainty in the period
634: from $\rho_{\star}$ is equal to the contribution from the {\it Kepler} photometry,
635: i.e.\ where the two terms in the brackets in Eq. \ref{eqn:p}
636: are equal. This occurs at a radius of,
637: \begin{equation}
638: r_{\rm p,crit} \simeq R_{\rm Nep} \left[10^{0.4(V-12)}
639: \left(\frac{ \sigma_{\rho_{\star}}/\rho_{\star}}{0.1}\right)^{-2}
640: \left(\frac{R_{\star}}{R_{\odot}}\right)^4
641: \left(\frac{M_{\star}}{M_{\odot}}\right)^{1/3}
642: \left(\frac{3.5 {\rm\, yrs}}{P}\right)^{1/3}\right]^{1/5},
643: \end{equation}
644: assuming $b=0$. Note that $r_{\rm p,crit}$ is a weak or extremely weak function of
645: all of the parameters except $R_{\star}$.
646:
647: Fig. \ref{fig:contour} illustrates this point by showing contours of
648: constant uncertainty in the period as a function of the radius and
649: period of the planet. The calculations are for systems with $R_{\star}=R_{\odot}$, $b=0.2$, and $V=12.0$. This figure shows that for
650: planets larger than the radius of Neptune, the uncertainty in the
651: period will be dominated by the uncertainty in the inferred stellar
652: density, whereas for planets much smaller than Neptune, an accurate
653: estimate of the period from the {\it Kepler} light curve will
654: be impossible.
655:
656: As shown above, the uncertainty in $P$ is dominated by the uncertainty in $\rho_{\star}$ and $\tau$. Our ability to determine $\tau$, and its uncertainty, depends both on its length and how many points we have during the ingress or egress. The length of $\tau$ depends on $v_{\rm tr,p}$ (which is a proxy for $P$ in the circular case), $b$, and $r_{\rm p}$. These properties are intrinsic to the system, but may be derived from the observables. We can explore how the contours shown in Fig. \ref{fig:contour} vary with impact parameter. Figure \ref{fig:impact} shows that a system with a larger impact parameter will have a smaller fractional uncertainty in $\tau$. Thus, a system with a larger impact parameter would have a smaller fractional uncertainty in $P$ at fixed period and planet radius.
657:
658: The other effect that influences the contours is the sampling of $\tau$. As the number of points taken during the ingress/egress becomes small, the fractional uncertainty in $\tau$, and hence in $P$, increases. In particular, if only one point is taken during the ingress, then the duration is relatively unconstrained. The probability of taking only one point during the ingress increases linearly from 0 to 1 as $\tau$ decreases from $2 dt$ to $dt$. In Fig. \ref{fig:deriv}, the Monte Carlo simulations for the fractional uncertainty in $P$ diverge from the simple model over this range. We indicate this region by the shaded portion of Fig. \ref{fig:contour}. Where $\tau=dt$, the fractional uncertainty in $P$ can be several times that predicted by the theoretical calculations, but these uncertainties converge as the shading gets lighter towards $\tau=2dt$. Below the line where $\tau=dt$, the uncertainty in $\tau$ increases rapidly, but fortunately, in much of this regime, {\it Kepler} will observe multiple transits, and this analysis will be unnecessary. As shown in the figure, a faster sampling rate, such as 1 per 20 min, significantly expands the parameter space over which our theoretical uncertainties are valid.
659:
660: \begin{figure}
661: \includegraphics[width=5.5in]{f6.eps}
662: \caption{
663: Uncertainty in derived quantities as a function of period. The
664: uncertainties in $r_{\rm p}$, $a$, $P$, $v_{\rm tr,p}$, and $b^2$ are plotted versus period for Neptune orbiting the Sun with an impact parameter of 0.2. The
665: line styles are the same as for Fig. \ref{fig:sigobs}. The vertical dotted lines mark where the ingress/egress duration $\tau$ is equal to the sampling rate $dt$, as well as where $\tau=2dt$ for 30-minute sampling.
666: \label{fig:deriv}}
667: \end{figure}
668:
669: \begin{figure}
670: \includegraphics[width=6in]{f7.eps}
671: \caption{
672: Contours of constant fractional uncertainty in the period as a function
673: of the period and the planet radius. The model is for a solar-type
674: star with a V magnitude of 12.0 and an impact parameter for the system
675: of 0.2. The dashed line indicates how the fractional uncertainty in $P$ changes with $V$ magnitude; it shows the 0.20 contour for $V=14.0$ and $b=0.2$. The diagonal black dotted line represents the boundary below which the
676: assumptions of Fisher matrix break down (the ingress/egress
677: time $\tau$ is roughly equal to the sampling rate of 1/30 min). Thus, the contours below this boundary are shown as dash-dotted lines. The region between $\tau=2dt$ and $\tau=dt$ is shaded to indicate the increasingly probability of obtaining only one point during ingress or egress leading to an uncertainty in $P$ that is larger than theoretical expectations. The diagonal gray dotted line indicates the $\tau=dt$ boundary for 20-min sampling. The vertical dotted
678: line shows the mission lifetime of {\it Kepler} ($L=3.5\, {\rm
679: yrs}$). Solar system planets are indicated. \label{fig:contour}}
680: \end{figure}
681:
682: \section{Estimating the Uncertainty in $m_{\rm p}$}
683: \label{sec:rv}
684: The mass of the planet comes from sampling the stellar radial velocity curve soon after the transit is observed.
685: Near the time of transit we can expand the stellar radial velocity,
686: \begin{equation}
687: \label{eqn:rv}
688: v_{\star} = v_0 - K_{\star}\sin\left[\frac{2\pi}{P}(t-t_c)\right]\approx v_0 - A_{\star}(t-t_c) ,
689: \end{equation}
690: where $v_0$ is the systemic velocity, $K_{\star}$ is the stellar radial
691: velocity semi-amplitude, and $A_{\star}\equiv 2\pi K_{\star}/P$ is the stellar
692: acceleration. Because the planet is known to transit and has a long
693: period, we assume $\sin i =1$. A Fisher matrix analysis of the
694: linear form of Eq. \ref{eqn:rv} gives the estimated uncertainty in
695: $A_{\star}$ to be
696: \begin{equation}
697: \sigma_{A_{\star}}^2 \simeq \frac{12\sigma_{\rm RV}^2}{T_{\rm tot}^2N},
698: \end{equation}
699: where $\sigma_{\rm RV}$ is the radial velocity precision, $T_{\rm
700: tot}$ is the total time span of the radial velocity observations, and
701: $N$ is the number of observations, and we have assumed $N \gg 1$ and
702: that the observations are evenly spaced in time. The details of this
703: derivation are given in the appendix ($\S$\ref{sec:append}).
704:
705: Equations for the mass of the planet and the uncertainty are
706: \begin{eqnarray}
707: m_{\rm p} &=& \frac{1}{16G}g_{\star}^2A_{\star}\left(\frac{T\tau}{\sqrt{\delta}}\right)^2\\
708: \sigma_{m_{\rm p}}^2 &=& m_{\rm p}^2\left[\frac{4}{g_{\star}^2}\sigma_{g_{\star}}^2+\frac{1}{A_{\star}^2}\sigma_{A_{\star}}^2+\frac{4}{T^2}\sigma_{T}^2+\frac{4}{\tau^2}\sigma_{\tau}^2+\frac{1}{\delta^2}\sigma_{\delta}^2+\frac{8}{T\tau}\sigma_{T\tau}^2-\frac{4}{T\delta}\sigma_{T\delta}^2-\frac{4}{\tau\delta}\sigma_{\tau\delta}^2\right]\nonumber\\
709: &\simeq& m_{\rm p}^2\left[\frac{4}{g_{\star}}\sigma_{g_{\star}}^2+\frac{1}{A_{\star}^2}\sigma_{A_{\star}}^2+Q^{-2}\left(\frac{24T}{\tau}\right)\right] . \nonumber
710: \end{eqnarray}
711: The approximate expression for the uncertainty in the mass is taken in
712: the limit $\tau \ll T \ll D$. Furthermore, the
713: $Q^{-2}\left(24T/\tau\right)$ term can be neglected in most cases
714: because, as we showed for the corresponding term in uncertainty in the
715: period, it will be very small for planets with radii larger than that
716: of Neptune. The scaling for the $\left(\sigma_{A_{\star}}/{A_{\star}}\right)^2$
717: term is given in Eq. \ref{eqn:sigBscale}.
718:
719: Contours of constant fractional uncertainty in the mass of the planet are shown
720: in Fig. \ref{fig:mpplot} as a function of $P$ and $m_{\rm p}$, assuming
721: $N=20$ radial velocity measurements are taken over $T_{\rm tot}=3$
722: mos with a precision $\sigma_{\rm RV} = 10 {\rm\, m\,s}^{-1}$. It shows that the planet mass can be estimated to
723: within a factor of two over this time period, thus
724: establishing the planetary nature of the transiting object.
725: By doubling the length of observations, one can put stronger constraints on
726: the mass of the planet.
727:
728: There are two points to bear in mind when applying this estimate for the
729: uncertainty in the mass of the planet. First, for $r_{\rm p} \lesssim R_{\rm Neptune}$, the $Q^{-2}(24T/\tau)$ term is no
730: longer small. The second consideration is that as time progresses away
731: from the time of transit, the straight line approximation to the
732: radial velocity curve will break down. In that case, the period will
733: begin to be constrained by the radial velocity curve itself, and the
734: uncertainties should be calculated from a Fisher matrix analysis of
735: the full expression for the radial velocity curve with three
736: parameters: $v_0$, $K_{\star}$, and $P$.
737:
738: \begin{figure}
739: \includegraphics[width=6in]{f8.eps}
740: \caption{
741: Contours of constant uncertainty in $m_{\rm p}$ as a function of $m_{\rm p}$ and
742: $P$. The solid lines show the result for 20 radial velocity measurements
743: with precision of $10~{\rm m~s^{-1}}$ taken
744: over a period of 3 months after the transit. The dashed line
745: shows the contour for $\sigma_{m_{\rm p}}/m_{\rm p} = 0.50$ for 40 observations taken over 6 months. The
746: dotted line indicates the mission lifetime of 3.5 years. The positions
747: of Jupiter and Saturn are indicated. \label{fig:mpplot}}
748: \end{figure}
749:
750: \section{Eccentricity}
751: \label{sec:e}
752:
753: The results presented in the previous sections assumed circular
754: orbits. Given that the average eccentricity $e$ of planets with $P
755: \ge 1~{\rm yr}$ is $e \simeq 0.3$, this
756: is not necessarily a good assumption. In this section, we assess the effects
757: of non-zero
758: eccentricities on the ability to characterize single-transit events
759: detected by {\it Kepler}. We note that our discussion has some
760: commonality with the study of \citet{Ford08}, who discuss the
761: possibility of characterizing the orbital eccentricities of
762: transiting planets with photometric observations. However, our
763: study addresses this topic from a very different perspective.
764:
765: In \S\ref{sec:basic} we demonstrated that, under the assumption that
766: $e=0$, the planet period $P$ and mass $m_{\rm p}$ can be inferred from two
767: observables: the velocity of the planet at the time of transit
768: $v_{\rm tr}$, and the projected acceleration of the star $A_{\star}$. In the
769: case of a non-zero eccentricity, there are two additional unknown
770: parameters: the eccentricity $e$ and the argument of periastron of the
771: planet $\omega_p$. Thus, with only two observables ($v_{\rm tr}$ and
772: $A_{\star}$) and four unknowns ($P, m_{\rm p}, e, \omega_p$), it is not possible
773: to obtain a unique solution.
774:
775: Although a unique solution to the planet parameters does not exist
776: using the observable parameters alone, it
777: may nevertheless be possible to obtain an interesting constraint on $P$ and/or
778: $m_{\rm p}$ by adopting reasonable priors on $e$ and $\omega_p$. From the transit
779: observables, one can estimate the velocity of the planet at the time of transit
780: (Eq. \ref{eqn:vtr}). In the case of $e\ne 0$, this is given by,
781: \begin{equation}
782: v_{\rm tr,p}=\frac{2\pi a}{P} \frac{ 1+e\sin \omega_p}{(1-e^2)^{1/2}},
783: \end{equation}
784: where here and throughout this section we assume $\sin i = 1$. Solving for $P$,
785: \begin{equation}
786: P=\frac{2\pi GM_{\star}}{v_{\rm tr,p}^3}\left[ \frac{1+e\sin \omega_p}{(1-e^2)^{1/2}}\right]^3.
787: \end{equation}
788: For a fixed value of $e$, the inferred period relative to the assumption
789: of a circular orbit has extremes due
790: to the unknown value of $\omega_p$ of
791: \begin{equation}
792: \left(\frac{\Delta P}{P}\right)_{min/max}=\left( \frac{1+e}{1-e}\right)^{\pm 3/2}.
793: \end{equation}
794: Taking a typical value for the eccentricity of $e=0.3$, this gives a range of
795: inferred periods of $0.4-2.5$ relative to the assumption of a circular orbit.
796: In fact, because the probability that a planet with a given $a$ transits
797: its parent star depends on $e$ and $\omega_p$ \citep{Barnes07,Burke08},
798: a proper Bayesian estimate, accounting for these selection effects,
799: might reduce the range of inferred values for $P$ significantly.
800:
801: What can be learned from RV observations immediately after the transit?
802: Expanding the stellar projected velocity around the time of transit, we can write,
803: \begin{equation}
804: v_{\star} = v_0 +e K_* \cos{\omega_p} - A_{\star}(t-t_c) + \frac{1}{2}J_{\star} (t-t_c)^2.
805: \end{equation}
806: The projected acceleration of the star at the time of transit is,
807: \begin{equation}
808: A_{\star} = \frac{Gm_{\rm p}}{a^2} \left( \frac{1+e\sin \omega_p}{1-e^2}\right)^{2}.
809: \end{equation}
810: This can be combined with $v_{\rm tr,p}$ to derive the mass of the planet,
811: \begin{equation}
812: m_{\rm p} = \frac{G M_{\star}^2}{v_{\rm tr,p}^4} A_{\star} (1+e \sin \omega_p)^2
813: \end{equation}
814: Thus, for a fixed eccentricity, the inferred mass relative to the assumption of a circular
815: orbit has extremes of,
816: \begin{equation}
817: \left(\frac{\Delta m_{\rm p}}{m_{\rm p}}\right)_{min/max}=(1\pm e)^2,
818: \end{equation}
819: which for $e=0.3$ yields a range of $0.5-1.7$, which is generally {\it smaller} than
820: the contribution to the uncertainty in $m_{\rm p}$ due to
821: the measurement uncertainty in $A_{\star}$ for our fiducial case of $m_{\rm p}=M_{\rm Jup}$ and $P=3.5~{\rm yrs}$ (see Eq. \ref{eqn:sigBscale}) .
822:
823: We may also consider what can be learned if it is possible measure the curvature of the
824: stellar radial velocity variations immediately after transit.
825: The projected stellar jerk is given by,
826: \begin{equation}
827: J_{\star} = \frac{4\pi G m_{\rm p}}{P a^2} \frac{ (1+e\sin\omega_p)^3 e \cos \omega_p}{(1-e^2)^{7/2}}.
828: \end{equation}
829: This can be combined with $v_{\rm tr}$ and $A_{\star}$ to provide an independent constraint on a combination
830: of the eccentricity and argument of periastron,
831: \begin{equation}
832: \frac{e\cos \omega_p}{(1+e\sin \omega_p)^2} = \frac{GM_{\star} J_{\star}}{2 A_{\star} v_{\rm tr}^3}.
833: \end{equation}
834: Unfortunately, it will be quite difficult to measure $J_{\star}$. Using the same Fisher formalism
835: as we used to estimate the uncertainty in $A_{\star}$ (\S\ref{sec:append}),
836: assuming $N$ evenly-spaced RV measurements with precision $\sigma_{\rm RV}$, taken
837: over time span $T_{\rm tot}$ after the transit, the uncertainty in $J_{\star}$ is,
838: \begin{equation}
839: \sigma_{J_{\star}}^2 \simeq \frac{720 \sigma_{\rm RV}^2}{N T_{\rm tot}^4},
840: \end{equation}
841: where we have assumed $N\gg 1$. For our fiducial parameters, the fractional
842: uncertainty in $J_{\star}$ is,
843: \begin{equation}
844: \left(\frac{\sigma_{J_{\star}}}{J_{\star}}\right)^2 \simeq
845: 710
846: \left(\frac{\sigma_{\rm RV}}{10 {\rm\, m\, s}^{-1}}\right)^2
847: \left(\frac{3 {\rm\, mos}}{T_{\rm tot}}\right)^4
848: \left(\frac{20}{N}\right)
849: \left(\frac{M_{\rm Jup}}{m_{\rm p}}\right)^2
850: \left(\frac{M_{\star}}{M_{\odot}}\right)^{4/3}
851: \left(\frac{P}{3.5 {\rm\, yrs}}\right)^{10/3}
852: \left(\frac{e\cos{\omega_p}}{0.3}\right)^{-2},
853: \end{equation}
854: where we have approximated $e \cos{\omega_p}(1+e\sin{\omega_p})^3(1-e^2)^{-7/2}\sim e\cos{\omega_p}$.
855: We conclude that, in order to obtain interesting constraints on
856: the eccentricity of long-period planets detected by {\it Kepler},
857: higher-precision RV measurements taken over a
858: baseline comparable to $\sim P$ will be needed.
859:
860:
861: \section{Summary}
862: \label{sec:conclude}
863:
864: The discovery of long-period transiting planets along with
865: subsequent follow-up observations would greatly enhance our understanding of the physics
866: of planetary atmospheres and interiors. Such planets would
867: allow us to gather constraints
868: in a regime of parameter space currently only occupied by our
869: own giant planets, namely planets whose energy budgets are dominated
870: by their residual internal heat, rather than by stellar insolation.
871: These constraints, in turn, might provide new insights into
872: planet formation. In this paper, we demonstrated that it will be possible to
873: detect and characterize such long-period planets using
874: observations of single transits by the
875: {\it Kepler} satellite, combined with precise radial velocity
876: measurements taken immediately after the transit. Indeed, these
877: results can be generalized to any transiting planet survey using the
878: scaling relations we provide, and it may be particularly interesting
879: to apply them to the {\it COnvection, ROtation \& planetary Transits
880: (CoRoT)} mission \citep{Baglin03}.
881:
882: We calculated that {\it Kepler} will see a few long-period, single
883: transit events and showed that, for circular orbits, the period of the system can be derived
884: from the {\it Kepler} light curve of a single event. We derived an
885: expression for the uncertainty in this period and showed that it is
886: dominated by the uncertainty in the stellar density derived from
887: spectroscopy (which we assume is $\sim 10\%$) for planets with radii
888: larger than the radius of Neptune, rather than being dependent on the
889: properties of the transit itself. This method can also be applied to
890: planets that {\it Kepler} will observe more than once, so that the
891: second time a planet is expected to transit {\it Kepler} can make a
892: selective improvement to its time sampling to better characterize the
893: transit.
894:
895: We have also shown how the mass of the planet can be constrained by
896: acquiring precise ($\sim 10~{\rm m~s^{-1}}$) radial velocity
897: measurements beginning shortly after the transit occurs. We have shown
898: that 20 measurements over 3 months can measure the mass of a
899: Jupiter-sized object to within a factor of a few and that extending
900: those observations to 40 measurements over 6 months significantly reduces
901: the uncertainty in the mass. Knowing the mass to within a factor of a
902: few in such a short time can distinguish between brown dwarfs and
903: planets rapidly and allow us to maximize our use of radial velocity
904: resources.
905:
906: We explored the effect of eccentricity on the ability to estimate the
907: planet mass and period. Allowing for a non-zero eccentricity adds two
908: additional parameters, and as a result it is not possible to obtain a
909: unique solution for the planet mass, period, eccentricity, and
910: argument of periastron. However, by adopting a reasonable prior on
911: the eccentricity of the planet, the period and mass of the planet can
912: still be estimated to within a factor of a few. Detailed
913: characterization of the planet properties will require precision RV
914: measurements obtained over a duration comparable to the period of the
915: planet.
916:
917: Thus, in the interest of ``getting the most for your money,'' we have
918: shown that the sensitivity of {\it Kepler} extends to planets with
919: periods beyond its nominal mission lifetime. With the launch of the
920: {\it Kepler} satellite, we are poised to discover and characterize
921: several long-period transiting systems, provided that we are prepared
922: to look for them.
923:
924: \acknowledgments We are grateful to Josh Carter, Jason Eastman, Eric Ford, Andy Gould, Matt Holman,
925: Yoram Lithwick, Josh Winn, and Andrew Youdin for helpful discussions. We thank the referee Ron Gilliland for constructive comments. JCY is supported by a Dean's Graduate Enrichment Fellowship from The Ohio State University.
926:
927: \bibliographystyle{apj}
928: \begin{thebibliography}{19}
929: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
930:
931: \bibitem[{{Baglin}(2003)}]{Baglin03}
932: {Baglin}, A. 2003, Advances in Space Research, 31, 345
933:
934: \bibitem[{{Barnes}(2007)}]{Barnes07}
935: {Barnes}, J.~W. 2007, \pasp, 119, 986
936:
937: \bibitem[{{Basri} {et~al.}(2005){Basri}, {Borucki}, \& {Koch}}]{Basri05}
938: {Basri}, G., {Borucki}, W.~J., \& {Koch}, D. 2005, New Astronomy Review, 49,
939: 478
940:
941: \bibitem[{{Beaulieu} {et~al.}(2006){Beaulieu}, {Bennett}, {Fouqu{\'e}},
942: {Williams}, {Dominik}, {J{\o}rgensen}, {Kubas}, {Cassan}, {Coutures},
943: {Greenhill}, {Hill}, {Menzies}, {Sackett}, {Albrow}, {Brillant}, {Caldwell},
944: {Calitz}, {Cook}, {Corrales}, {Desort}, {Dieters}, {Dominis}, {Donatowicz},
945: {Hoffman}, {Kane}, {Marquette}, {Martin}, {Meintjes}, {Pollard}, {Sahu},
946: {Vinter}, {Wambsganss}, {Woller}, {Horne}, {Steele}, {Bramich}, {Burgdorf},
947: {Snodgrass}, {Bode}, {Udalski}, {Szyma{\'n}ski}, {Kubiak}, {Wi{\c e}ckowski},
948: {Pietrzy{\'n}ski}, {Soszy{\'n}ski}, {Szewczyk}, {Wyrzykowski},
949: {Paczy{\'n}ski}, {Abe}, {Bond}, {Britton}, {Gilmore}, {Hearnshaw}, {Itow},
950: {Kamiya}, {Kilmartin}, {Korpela}, {Masuda}, {Matsubara}, {Motomura},
951: {Muraki}, {Nakamura}, {Okada}, {Ohnishi}, {Rattenbury}, {Sako}, {Sato},
952: {Sasaki}, {Sekiguchi}, {Sullivan}, {Tristram}, {Yock}, \&
953: {Yoshioka}}]{Beaulieu06}
954: {Beaulieu}, J.-P., {Bennett}, D.~P., {Fouqu{\'e}}, P., {Williams}, A.,
955: {Dominik}, M., {J{\o}rgensen}, U.~G., {Kubas}, D., {Cassan}, A., {Coutures},
956: C., {Greenhill}, J., {Hill}, K., {Menzies}, J., {Sackett}, P.~D., {Albrow},
957: M., {Brillant}, S., {Caldwell}, J.~A.~R., {Calitz}, J.~J., {Cook}, K.~H.,
958: {Corrales}, E., {Desort}, M., {Dieters}, S., {Dominis}, D., {Donatowicz}, J.,
959: {Hoffman}, M., {Kane}, S., {Marquette}, J.-B., {Martin}, R., {Meintjes}, P.,
960: {Pollard}, K., {Sahu}, K., {Vinter}, C., {Wambsganss}, J., {Woller}, K.,
961: {Horne}, K., {Steele}, I., {Bramich}, D.~M., {Burgdorf}, M., {Snodgrass}, C.,
962: {Bode}, M., {Udalski}, A., {Szyma{\'n}ski}, M.~K., {Kubiak}, M., {Wi{\c
963: e}ckowski}, T., {Pietrzy{\'n}ski}, G., {Soszy{\'n}ski}, I., {Szewczyk}, O.,
964: {Wyrzykowski}, {\L}., {Paczy{\'n}ski}, B., {Abe}, F., {Bond}, I.~A.,
965: {Britton}, T.~R., {Gilmore}, A.~C., {Hearnshaw}, J.~B., {Itow}, Y., {Kamiya},
966: K., {Kilmartin}, P.~M., {Korpela}, A.~V., {Masuda}, K., {Matsubara}, Y.,
967: {Motomura}, M., {Muraki}, Y., {Nakamura}, S., {Okada}, C., {Ohnishi}, K.,
968: {Rattenbury}, N.~J., {Sako}, T., {Sato}, S., {Sasaki}, M., {Sekiguchi}, T.,
969: {Sullivan}, D.~J., {Tristram}, P.~J., {Yock}, P.~C.~M., \& {Yoshioka}, T.
970: 2006, \nat, 439, 437
971:
972: \bibitem[{{Borucki} {et~al.}(2004){Borucki}, {Koch}, {Boss}, {Dunham},
973: {Dupree}, {Geary}, {Gilliland}, {Howell}, {Jenkins}, {Kondo}, {Latham},
974: {Lissauer}, \& {Reitsema}}]{Borucki04}
975: {Borucki}, W., {Koch}, D., {Boss}, A., {Dunham}, E., {Dupree}, A., {Geary}, J.,
976: {Gilliland}, R., {Howell}, S., {Jenkins}, J., {Kondo}, Y., {Latham}, D.,
977: {Lissauer}, J., \& {Reitsema}, H. 2004, in ESA Special Publication, Vol. 538,
978: Stellar Structure and Habitable Planet Finding, ed. F.~{Favata},
979: S.~{Aigrain}, \& A.~{Wilson}, 177--182
980:
981: \bibitem[{{Burke}(2008)}]{Burke08}
982: {Burke}, C.~J. 2008, arXiv:0801.2579, 801
983:
984: \bibitem[{{Butler} {et~al.}(2006){Butler}, {Wright}, {Marcy}, {Fischer},
985: {Vogt}, {Tinney}, {Jones}, {Carter}, {Johnson}, {McCarthy}, \&
986: {Penny}}]{Butler06}
987: {Butler}, R.~P., {Wright}, J.~T., {Marcy}, G.~W., {Fischer}, D.~A., {Vogt},
988: S.~S., {Tinney}, C.~G., {Jones}, H.~R.~A., {Carter}, B.~D., {Johnson}, J.~A.,
989: {McCarthy}, C., \& {Penny}, A.~J. 2006, \apj, 646, 505
990:
991: \bibitem[{{Carter} {et~al.}(2008){Carter}, {Yee}, {Eastman}, {Gaudi}, \&
992: {Winn}}]{Carter08}
993: {Carter}, J.~A., {Yee}, J.~C., {Eastman}, J., {Gaudi}, B.~S., \& {Winn}, J.~N.
994: 2008, arXiv:0805.0238, 805
995:
996: \bibitem[{{Charbonneau} {et~al.}(2005){Charbonneau}, {Allen}, {Megeath},
997: {Torres}, {Alonso}, {Brown}, {Gilliland}, {Latham}, {Mandushev}, {O'Donovan},
998: \& {Sozzetti}}]{Charbonneau05}
999: {Charbonneau}, D., {Allen}, L.~E., {Megeath}, S.~T., {Torres}, G., {Alonso},
1000: R., {Brown}, T.~M., {Gilliland}, R.~L., {Latham}, D.~W., {Mandushev}, G.,
1001: {O'Donovan}, F.~T., \& {Sozzetti}, A. 2005, \apj, 626, 523
1002:
1003: \bibitem[{{Charbonneau} {et~al.}(2002){Charbonneau}, {Brown}, {Noyes}, \&
1004: {Gilliland}}]{Charbonneau02}
1005: {Charbonneau}, D., {Brown}, T.~M., {Noyes}, R.~W., \& {Gilliland}, R.~L. 2002,
1006: \apj, 568, 377
1007:
1008: \bibitem[{{Claret}(2000)}]{Claret00}
1009: {Claret}, A. 2000, \aap, 363, 1081
1010:
1011: \bibitem[{{Cumming} {et~al.}(2008){Cumming}, {Butler}, {Marcy}, {Vogt},
1012: {Wright}, \& {Fischer}}]{Cumming08}
1013: {Cumming}, A., {Butler}, R.~P., {Marcy}, G.~W., {Vogt}, S.~S., {Wright}, J.~T.,
1014: \& {Fischer}, D.~A. 2008, arXiv:0803.3357, 803
1015:
1016: \bibitem[{{Deming} {et~al.}(2005){Deming}, {Seager}, {Richardson}, \&
1017: {Harrington}}]{Deming05}
1018: {Deming}, D., {Seager}, S., {Richardson}, L.~J., \& {Harrington}, J. 2005,
1019: \nat, 434, 740
1020:
1021: \bibitem[{{Ford} {et~al.}(2008){Ford}, {Quinn}, \& {Veras}}]{Ford08}
1022: {Ford}, E.~B., {Quinn}, S.~N., \& {Veras}, D. 2008, arXiv:0801.2591, 801
1023:
1024: \bibitem[{{Gould}(2003)}]{Gould03}
1025: {Gould}, A. 2003, arXiv:astro-ph/0310577
1026:
1027: \bibitem[{{Gould} {et~al.}(2006){Gould}, {Udalski}, {An}, {Bennett}, {Zhou},
1028: {Dong}, {Rattenbury}, {Gaudi}, {Yock}, {Bond}, {Christie}, {Horne},
1029: {Anderson}, {Stanek}, {DePoy}, {Han}, {McCormick}, {Park}, {Pogge},
1030: {Poindexter}, {Soszy{\'n}ski}, {Szyma{\'n}ski}, {Kubiak}, {Pietrzy{\'n}ski},
1031: {Szewczyk}, {Wyrzykowski}, {Ulaczyk}, {Paczy{\'n}ski}, {Bramich},
1032: {Snodgrass}, {Steele}, {Burgdorf}, {Bode}, {Botzler}, {Mao}, \&
1033: {Swaving}}]{Gould06}
1034: {Gould}, A., {Udalski}, A., {An}, D., {Bennett}, D.~P., {Zhou}, A.-Y., {Dong},
1035: S., {Rattenbury}, N.~J., {Gaudi}, B.~S., {Yock}, P.~C.~M., {Bond}, I.~A.,
1036: {Christie}, G.~W., {Horne}, K., {Anderson}, J., {Stanek}, K.~Z., {DePoy},
1037: D.~L., {Han}, C., {McCormick}, J., {Park}, B.-G., {Pogge}, R.~W.,
1038: {Poindexter}, S.~D., {Soszy{\'n}ski}, I., {Szyma{\'n}ski}, M.~K., {Kubiak},
1039: M., {Pietrzy{\'n}ski}, G., {Szewczyk}, O., {Wyrzykowski}, {\L}., {Ulaczyk},
1040: K., {Paczy{\'n}ski}, B., {Bramich}, D.~M., {Snodgrass}, C., {Steele}, I.~A.,
1041: {Burgdorf}, M.~J., {Bode}, M.~F., {Botzler}, C.~S., {Mao}, S., \& {Swaving},
1042: S.~C. 2006, \apjl, 644, L37
1043:
1044: \bibitem[{{Mandel} \& {Agol}(2002)}]{Mandel02}
1045: {Mandel}, K. \& {Agol}, E. 2002, \apjl, 580, L171
1046:
1047: \bibitem[{{Rowe} {et~al.}(2007){Rowe}, {Matthews}, {Seager}, {Miller-Ricci},
1048: {Sasselov}, {Kuschnig}, {Guenther}, {Moffat}, {Rucinski}, {Walker}, \&
1049: {Weiss}}]{Rowe07}
1050: {Rowe}, J.~F., {Matthews}, J.~M., {Seager}, S., {Miller-Ricci}, E., {Sasselov},
1051: D., {Kuschnig}, R., {Guenther}, D.~B., {Moffat}, A.~F.~J., {Rucinski}, S.~M.,
1052: {Walker}, G.~A.~H., \& {Weiss}, W.~W. 2007, arXiv:0711.4111, 711
1053:
1054: \bibitem[{{Seager} \& {Mall{\'e}n-Ornelas}(2003)}]{Seager03}
1055: {Seager}, S. \& {Mall{\'e}n-Ornelas}, G. 2003, \apj, 585, 1038
1056:
1057: \end{thebibliography}
1058:
1059:
1060: \section{Appendix: Detailed Derivations}
1061: \label{sec:append}
1062:
1063: \subsection{Derivation of the Uncertainties in the Light Curve Observables}
1064:
1065: The Fisher matrix formalism is a simple way to estimate the
1066: uncertainties in the parameters, $\alpha$, of a model $F(x)$, that is being
1067: fit to a series of measurements $x_k$ with measurement uncertainties $\sigma_k$. The
1068: covariance of $\alpha_i$ with $\alpha_j$ is given by the element $c_{ij}$ of the covariance matrix
1069: $c$, where $c = b^{-1}$ and the entries of $b$ are given by,
1070: \begin{equation}
1071: b_{ij} \equiv \displaystyle\sum_{k=1}^{N_d} \frac{\partial F(x_k)}{\partial \alpha_i}\frac{\partial F(x_k)}{\partial \alpha_j} \frac{1}{\sigma_k^2},
1072: \label{bij}
1073: \end{equation}
1074: where $N_d$ is the number of data points. In the limit of infinite sampling ($N_d \rightarrow \infty$) and fixed precision, $\sigma_k=\sigma$,
1075: \begin{equation}
1076: b_{ij} \rightarrow \frac{1}{D\sigma^2} \displaystyle\int_{0}^{D} \frac{\partial F(x)}{\partial \alpha_i}\frac{\partial F(x)}{\partial \alpha_j}\,dx,
1077: \end{equation}
1078: where the interval of interest is given by $x = [0, D]$, and in this
1079: case $D$ is the total duration of observations. Thus, if the partial
1080: derivatives of a model with respect to its parameters are known, then
1081: the uncertainties in those parameters can be estimated
1082: \citep{Gould03}.
1083:
1084: For the simplified transit model described in $\S$\ref{sec:models},
1085: the observable parameters are $\alpha = [t_c, T, \tau, \delta, F_0]$. For
1086: a sampling rate $\gamma$, the $b$ matrix is,
1087: \begin{equation}
1088: b = \frac{\gamma}{\sigma^2}\left(\begin{array}{ccccc}
1089: \frac{2\delta^2}{\tau} & 0 & 0 & 0 & 0\\
1090: 0 & \frac{\delta^2}{2\tau} & 0 & \frac{\delta}{2} & -\delta\\
1091: 0 & 0 & \frac{\delta^2}{6\tau} & -\frac{\delta}{6} & 0\\
1092: 0 & \frac{\delta}{2} & -\frac{\delta}{6} & T-\frac{\tau}{3} & -T\\
1093: 0 & -\delta & 0 & -T & D
1094: \end{array}\right).
1095: \end{equation}
1096: The covariance matrix is,
1097: \begin{equation}
1098: c = b^{-1} = \left(\frac{\sigma^2}{\gamma T\delta^2}\right)
1099: \left(\begin{array}{ccccc}
1100: \frac{\tau T}{2} & 0 & 0 & 0 & 0\\
1101: 0 & -\frac{\tau T(D\tau-2DT+2T^2)}{t_{out}t_{f}} &
1102: -\frac{\tau^2T(D-2T)}{t_{out}t_{f}} &
1103: -\frac{\tau T\delta(D-2T)}{t_{out}t_{f}} & \frac{\tau T\delta}{t_{out}}\\
1104: 0 & -\frac{\tau^2T(D-2T)}{t_{out}t_{f}} &
1105: -\frac{\tau T(5D\tau-4\tau^2-6DT+6T^2)}{t_{out}t_{f}} &
1106: \frac{\tau T\delta(D-2\tau)}{t_{out}t_{f}} & \frac{\tau T\delta}{t_{out}}\\
1107: 0 & -\frac{\tau T\delta(D-2T)}{t_{out}t_{f}} &
1108: \frac{\tau T\delta(D-2\tau)}{t_{out}t_{f}} & \frac{T\delta^2(D-2\tau)}{t_{out}t_f} &
1109: \frac{T\delta^2}{t_{out}}\\
1110: 0 & \frac{\tau T\delta}{t_{out}} & \frac{\tau T\delta}{t_{out}} &
1111: \frac{T\delta^2}{t_{out}} & \frac{T\delta^2}{t_{out}}
1112: \end{array}\right),
1113: \end{equation}
1114: where $t_{f} \equiv T-\tau$ is the duration of the flat part of the
1115: eclipse and $t_{out} \equiv D-\tau-T$ is the time spent out of
1116: eclipse. Thus, the uncertainty on the ingress/egress time, $\sigma_{\tau}$,
1117: is given by
1118: \begin{equation}
1119: \sigma_{\tau} = \sqrt{c_{33}}=\frac{\sigma}{\sqrt{\gamma}}
1120: \sqrt{\frac{-\tau(5D\tau-4\tau^2-6DT+6T^2)}{\delta^2t_{out}t_{f}}}.
1121: \end{equation}
1122:
1123: Define $Q \equiv \sqrt{\gamma T}(\delta/\sigma)$. Then,
1124: in the limit $\tau \ll T$, the uncertainties in the observable parameters are given by,
1125: \begin{eqnarray}
1126: & \sigma_{t_c} & = Q^{-1}\sqrt{\frac{T\tau}{2}}\label{tc2}\\
1127: & \frac{\sigma_T}{T} & = Q^{-1}\sqrt{\frac{2\tau}{T}}\\
1128: & \frac{\sigma_{\tau}}{\tau} & = Q^{-1}\sqrt{\frac{6T}{\tau}}\\
1129: & \frac{\sigma_{\delta}}{\delta} & = Q^{-1}\sqrt{\frac{1}{\left(1-\frac{T}{D}\right)}}\\
1130: & \frac{\sigma_{F_0}}{F_0} & = 0 \mbox{ , since in general, $\delta \ll F_0$}.
1131: \label{fo2}
1132: \end{eqnarray}
1133: Note that $Q$ is approximately the total signal-to-noise ratio of the transit.
1134: Assuming that the photometric uncertainties are limited by photon noise,
1135: we have that $\gamma/\sigma^2 = \Gamma_{\rm ph}$, where $\Gamma_{\rm ph}$ is the photon collection rate.
1136: This recovers the expression for $Q$ in \S\ref{sec:basic}.
1137:
1138: A more detailed analysis of the variances and covariances of the
1139: transit observables can be found in \citet{Carter08}.
1140:
1141: \subsection{Derivation of the Uncertainties from the Radial Velocity Curve}
1142: A similar analysis can be done for the radial velocity curve, except
1143: that we use Eq. \ref{bij} and consider discrete observations. The radial
1144: velocity for a circular orbit is
1145:
1146: \begin{equation}
1147: v_{\star} = v_0 - K_{\star}\sin\left[\frac{2\pi}{P} (t-t_c)\right]
1148: \end{equation}
1149: Expanding about $t_c$ gives
1150: \begin{equation}
1151: v_{\star} = v_0 -K_{\star} \frac{2\pi}{P} (t-t_c) = v_0 - A_{\star}(t-t_c),
1152: \label{eqn:vexpand}
1153: \end{equation}
1154: where $A_{\star}$ is the stellar projected acceleration at the time of the transit,
1155: \begin{equation}
1156: A_{\star} = \frac{2\pi K_{\star}}{P} .
1157: \end{equation}
1158:
1159: Consider $N$ measurements of $v_{\star}$, each
1160: with precision $\sigma_{\rm RV}$, taken at times $t_k$. Fitting these measurements
1161: to the linear model in Eq. \ref{eqn:vexpand}, we can estimate
1162: the uncertainties in the parameters $v_0$ and $A_{\star}$ using the Fisher matrix
1163: formalism,
1164: \begin{equation}
1165: b = \frac{1}{\sigma^2}\left[\begin{array}{cc}
1166: N & -\displaystyle\sum_{k=1}^{N}(t_k-t_c)\\
1167: -\displaystyle\sum_{k=1}^{N}(t_k-t_c) & \displaystyle\sum_{k=1}^{N}(t_k-t_c)^2
1168: \end{array}\right] .
1169: \end{equation}
1170: The covariance matrix is the inverse of $b$,
1171: \begin{equation}
1172: c = \frac{\sigma^2}{\left[\displaystyle\sum_{k=1}^{N}(t_k-t_c)^2\right]-\left[\displaystyle\sum_{k=1}^{N}(t_k-t_c)\right]^2}\left[\begin{array}{cc}
1173: \displaystyle\sum_{k=1}^{N}(t_k-t_c)^2 & \displaystyle\sum_{k=1}^{N}(t_k-t_c)\\
1174: \displaystyle\sum_{k=1}^{N}(t_k-t_c) & N
1175: \end{array}\right] .
1176: \end{equation}
1177: If the points are evenly spaced by $\Delta t$
1178: \begin{eqnarray}
1179: && \displaystyle\sum_{k=1}^{N}(t_k-t_c) = \displaystyle\sum_{k=1}^{N} k\Delta t = \frac{N(N+1)}{2}\Delta t , \\
1180: && \displaystyle\sum_{k=1}^{N}(t_k-t_c)^2 = \displaystyle\sum_{k=1}^{N} (k\Delta t)^2 = \frac{N(N+1)(2N+1)}{6}\Delta t^2 ,
1181: \end{eqnarray}
1182: and thus,
1183: \begin{eqnarray}
1184: c &=& \frac{\sigma^2}{N(N-1)\Delta t^2}\left[\begin{array}{cc}
1185: 2(2N+1)\Delta t^2 & 6\Delta t\\
1186: 6 \Delta t & \frac{12}{(N+1)}
1187: \end{array}\right] .
1188: \end{eqnarray}
1189: The uncertainty in the projected stellar acceleration is,
1190: \begin{equation}
1191: \sigma_{A_{\star}}^2 = \frac{\sigma_{\rm RV}^2}{N(N-1)(N+1)(\Delta t)^2} .
1192: \end{equation}
1193:
1194: In the limit as $N \rightarrow \infty$, the covariance matrix reduces to,
1195: \begin{eqnarray}
1196: c &=& \frac{\sigma_{\rm RV}^2}{\Delta t^2}\left[\begin{array}{cc}
1197: \frac{4\Delta t^2}{N} & \frac{6\Delta t}{N^2}\\
1198: \frac{6 \Delta t}{N^2} & \frac{12}{N^3}
1199: \end{array}\right] .
1200: \end{eqnarray}
1201: Defining the total length of observations as $T_{\rm tot} \equiv N \Delta t$, we can also write,
1202: \begin{eqnarray}
1203: c &=& \frac{\sigma_{\rm RV}^2}{T_{\rm tot}^2}\left[\begin{array}{cc}
1204: 4\frac{T_{\rm tot}}{N} & 6\frac{T_{\rm tot}}{N}\\
1205: 6 \frac{T_{\rm tot}}{N} & \frac{12}{N}
1206: \end{array}\right],
1207: \end{eqnarray}
1208: and so the uncertainty $A_{\star}$ in the limit of $N \rightarrow \infty$ becomes,
1209: \begin{equation}
1210: \sigma_{A_{\star}}^2 = \frac{12 \sigma_{\rm RV}^2}{T_{\rm tot}^2 N}.
1211: \end{equation}
1212:
1213:
1214: \end{document}
1215: