transit timing variations for planets near

13
TRANSIT TIMING VARIATIONS FOR PLANETS NEAR ECCENTRICITY-TYPE MEAN MOTION RESONANCES Katherine M. Deck 1 and Eric Agol 2,3 1 Department of Geological and Planetary Sciences, California Institute of Technology, Pasadena, CA, USA; [email protected] 2 Department of Astronomy, University of Washington, Seattle, WA, USA 3 NASA Astrobiology Institutes Virtual Planetary Laboratory, Seattle, WA 98195, USA Received 2015 September 28; accepted 2016 January 20; published 2016 April 18 ABSTRACT We derive the transit timing variations (TTVs) of two planets near a second-order mean motion resonance (MMR) on nearly circular orbits. We show that the TTVs of each planet are given by sinusoids with a frequency of - - jn j n 2 2 1 ( ) , where j 3 is an integer characterizing the resonance and n 2 and n 1 are the mean motions of the outer and inner planets, respectively. The amplitude of the TTV depends on the mass of the perturbing planet, relative to the mass of the star, and on both the eccentricities and longitudes of pericenter of each planet. The TTVs of the two planets are approximated anti-correlated, with phases of f and f p » + , where the phase f also depends on the eccentricities and longitudes of pericenter. Therefore, the TTVs caused by proximity to a second-order MMR do not in general uniquely determine both planet masses, eccentricities, and pericenters. This is completely analogous to the case of TTVs induced by two planets near a rst-order MMR. We explore how other TTV signals, such as the short-period synodic TTV or a rst-order resonant TTV, in combination with the second-order resonant TTV, can break degeneracies. Finally, we derive approximate formulae for the TTVs of planets near any order eccentricity-type MMR; this shows that the same basic sinusoidal TTV structure holds for all eccentricity-type resonances. Our general formula reduces to previously derived results near rst-order MMRs. Key words: celestial mechanics planets and satellites: dynamical evolution and stability planets and satellites: fundamental parameters 1. INTRODUCTION Transit timing variations (TTVs; Miralda-Escudé 2002; Agol et al. 2005; Holman & Murray 2005) have proved useful for constraining the masses and orbital elements of exoplanets (e.g., Carter et al. 2012; Huber et al. 2013; Nesvorný et al. 2013). To date, most TTV studies have focused on pairs of planets near mean motion resonances (MMRs) and in particular on those near rst-order resonances. This is because near resonances the small perturbations that planets impart on each other can add coherently over time to produce a large, detectable TTV signal, and because for low eccentricity orbits the near resonance regionis largest for rst-order resonances. However, TTV models for planet pairs near resonance are plagued by degeneracies. For a pair of planets on coplanar orbits, there are 10 free parameters that control the TTVs. Assuming both planets transit, the periods and initial phases of the orbits are well-known, leaving six unknown parametersthe masses, eccentricities, and longitudes of pericenters of the two planets. Boué et al. (2012) showed analytically that all of these unknown parameters affect TTV amplitudes and phases for systems near or in eccentricity-type MMRs and they concluded that degeneracies between mass and orbital para- meters could strongly affect inferences based on the TTVs. The particular case of a pair of planets near (but not in) rst-order resonances was analyzed in detail by Lithwick et al. (2012), who showed that the TTVs of each planet are approximately sinusoidal, with a period set by the known mean orbital periods. Furthermore, the rst-order resonant TTVs are often nearly anti-correlated, in which case there are only three constraining observables: two TTV amplitudes and a single phase. Therefore, the six unknown parameters cannot all be determined uniquely, and in particular, a degeneracy between the masses and eccentricities results (Lithwick et al. 2012). More recently it has been demonstrated that a small amplitude choppingsignal associated with individual planetary conjunctionsand not with proximity to the MMR can determine the masses of the interacting planets uniquely, with only weak dependence on the eccentricities and longitudes of pericenter (Nesvorný & Vokrouhlický 2014; Deck & Agol 2015). If this chopping TTV is measured for a system near a rst-order resonance, the amplitude and phase of the resonant TTV can be used to constrain the remaining degrees of freedom (the eccentricities e and the longitudes of pericenter ϖ). However, as shown by Lithwick et al. (2012), the individual eccentricities and longitudes of pericenter are not constrained by the rst-order resonant TTV; rather, only a linear combination of the eccentricity vectors v v e e cos , sin ( ) are (the quantity Z free in the notation of Lithwick et al. 2012). This raises the question of if and in which circumstances TTVs can be used to measure individual eccentricities uniquely. Here we consider the case of two planets orbiting near a second-order resonance, with » - / P P j j 2 2 1 ( ) where j 3. Second-order resonances distinct from rst-order commensur- abilities appear when j is odd. Though less common than rst- order MMRs in the observed sample of transiting planets (Fabrycky et al. 2014), these congurations are still of interest and important to understand (Jontof-Hutter et al. 2015; Petigura et al. 2015). On the other hand, second-order resonances when j is even are also important because they represent the dominant correction at O e 2 ( ) to the rst-order resonant TTV formula derived by Lithwick et al. (2012). Because of the different functional dependence on the eccentricities and longitudes of pericenter between the rst- and second-order terms, second- The Astrophysical Journal, 821:96 (13pp), 2016 April 20 doi:10.3847/0004-637X/821/2/96 © 2016. The American Astronomical Society. All rights reserved. 1

Upload: doankhuong

Post on 04-Jan-2017

230 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: TRANSIT TIMING VARIATIONS FOR PLANETS NEAR

TRANSIT TIMING VARIATIONS FOR PLANETS NEAR ECCENTRICITY-TYPEMEAN MOTION RESONANCES

Katherine M. Deck1and Eric Agol

2,3

1 Department of Geological and Planetary Sciences, California Institute of Technology, Pasadena, CA, USA; [email protected] Department of Astronomy, University of Washington, Seattle, WA, USA

3 NASA Astrobiology Institutes Virtual Planetary Laboratory, Seattle, WA 98195, USAReceived 2015 September 28; accepted 2016 January 20; published 2016 April 18

ABSTRACT

We derive the transit timing variations (TTVs) of two planets near a second-order mean motion resonance (MMR)on nearly circular orbits. We show that the TTVs of each planet are given by sinusoids with a frequency of

- -jn j n22 1( ) , where j 3 is an integer characterizing the resonance and n2 and n1 are the mean motions of theouter and inner planets, respectively. The amplitude of the TTV depends on the mass of the perturbing planet,relative to the mass of the star, and on both the eccentricities and longitudes of pericenter of each planet. The TTVsof the two planets are approximated anti-correlated, with phases of f and f p» + , where the phase f also dependson the eccentricities and longitudes of pericenter. Therefore, the TTVs caused by proximity to a second-orderMMR do not in general uniquely determine both planet masses, eccentricities, and pericenters. This is completelyanalogous to the case of TTVs induced by two planets near a first-order MMR. We explore how other TTV signals,such as the short-period synodic TTV or a first-order resonant TTV, in combination with the second-order resonantTTV, can break degeneracies. Finally, we derive approximate formulae for the TTVs of planets near any ordereccentricity-type MMR; this shows that the same basic sinusoidal TTV structure holds for all eccentricity-typeresonances. Our general formula reduces to previously derived results near first-order MMRs.

Key words: celestial mechanics – planets and satellites: dynamical evolution and stability – planets and satellites:fundamental parameters

1. INTRODUCTION

Transit timing variations (TTVs; Miralda-Escudé 2002; Agolet al. 2005; Holman & Murray 2005) have proved useful forconstraining the masses and orbital elements of exoplanets(e.g., Carter et al. 2012; Huber et al. 2013; Nesvorný et al.2013). To date, most TTV studies have focused on pairs ofplanets near mean motion resonances (MMRs) and in particularon those near first-order resonances. This is because nearresonances the small perturbations that planets impart on eachother can add coherently over time to produce a large,detectable TTV signal, and because for low eccentricityorbits the “near resonance region” is largest for first-orderresonances.

However, TTV models for planet pairs near resonance areplagued by degeneracies. For a pair of planets on coplanarorbits, there are 10 free parameters that control the TTVs.Assuming both planets transit, the periods and initial phases ofthe orbits are well-known, leaving six unknown parameters—the masses, eccentricities, and longitudes of pericenters of thetwo planets. Boué et al. (2012) showed analytically that all ofthese unknown parameters affect TTV amplitudes and phasesfor systems near or in eccentricity-type MMRs and theyconcluded that degeneracies between mass and orbital para-meters could strongly affect inferences based on the TTVs. Theparticular case of a pair of planets near (but not in) first-orderresonances was analyzed in detail by Lithwick et al. (2012),who showed that the TTVs of each planet are approximatelysinusoidal, with a period set by the known mean orbitalperiods. Furthermore, the first-order resonant TTVs are oftennearly anti-correlated, in which case there are only threeconstraining “observables”: two TTV amplitudes and a singlephase. Therefore, the six unknown parameters cannot all be

determined uniquely, and in particular, a degeneracy betweenthe masses and eccentricities results (Lithwick et al. 2012).More recently it has been demonstrated that a small

amplitude “chopping” signal associated with individualplanetary conjunctions—and not with proximity to the MMR—can determine the masses of the interacting planets uniquely,with only weak dependence on the eccentricities and longitudesof pericenter (Nesvorný & Vokrouhlický 2014; Deck &Agol 2015). If this chopping TTV is measured for a systemnear a first-order resonance, the amplitude and phase of theresonant TTV can be used to constrain the remaining degreesof freedom (the eccentricities e and the longitudes of pericenterϖ). However, as shown by Lithwick et al. (2012), theindividual eccentricities and longitudes of pericenter are notconstrained by the first-order resonant TTV; rather, only alinear combination of the eccentricity vectors v ve ecos , sin( )are (the quantity Zfree in the notation of Lithwick et al. 2012).This raises the question of if and in which circumstances TTVscan be used to measure individual eccentricities uniquely.Here we consider the case of two planets orbiting near a

second-order resonance, with » -/P P j j 22 1 ( ) where j 3.Second-order resonances distinct from first-order commensur-abilities appear when j is odd. Though less common than first-order MMRs in the observed sample of transiting planets(Fabrycky et al. 2014), these configurations are still of interestand important to understand (Jontof-Hutter et al. 2015; Petiguraet al. 2015). On the other hand, second-order resonances when jis even are also important because they represent the dominantcorrection at O e2( ) to the first-order resonant TTV formuladerived by Lithwick et al. (2012). Because of the differentfunctional dependence on the eccentricities and longitudes ofpericenter between the first- and second-order terms, second-

The Astrophysical Journal, 821:96 (13pp), 2016 April 20 doi:10.3847/0004-637X/821/2/96© 2016. The American Astronomical Society. All rights reserved.

1

Page 2: TRANSIT TIMING VARIATIONS FOR PLANETS NEAR

order effects may be important for breaking degeneraciespresent in the TTVs of planets near first-order MMR.

We derive an approximate formula for TTVs resulting froman orbital configuration near a second-order resonance inSection 2. In Section 3, we interpret the resulting orbitalparameter and mass constraints allowed by these TTVs. InSection 4 we test the TTV formulae in multiple ways. We firstcompare the predicted TTVs using the formulae with thosedetermined via numerical integration of the full gravitationalequations of motion across a wide range of relevant parameterspace. Then, for the specific systems Kepler-26 and Kepler-46,we compare the outcome of TTV inversions obtained using theformulae to those obtained via n-body analysis. In Section 4.2.3we apply our formula to simulated data and investigate ifmeasuring the second harmonic of the TTVs for a pair ofplanets near a first-order resonance allows unique determina-tions of both planet eccentricities. For completeness, we extendour derivation to systems near the j:j−N Nth order MMR inSection 5. We give our conclusions in Section 6.

2. DERIVATION OF THE APPROXIMATE TTV

We would like to determine approximate expressions for theTTVs induced for two planets near the j:j−2 second-orderresonance. To begin, we write the Hamiltonian in Jacobielements up to the second-order in planet eccentricities. Weinclude only the second-order resonant terms in this derivation.In this case, the Hamiltonian is

a q v a

q v a q v v

=- - -

´ - +

´ - + - -

HGM m

a

GM m

a

Gm m

a

g e g e

g e e

2 2

cos 2

cos 2 cos

1

j j j

j j j

1

1

2

2

1 2

2

,45 12

1 ,53 22

2 ,49 1 2 2 1

[ ( ) ( ) ( )( ) ( ) ( )]

( )

where

q l l= - -j j 2 2j 2 1( ) ( )

and

ò

a a a

a

a a a

a

a a a

aad

aa

ap

q

a q aq

= - + + -

+

= - + - + -

-

= - + + -

+ -

º

=- +

a

a

a

a

a

a

a

p

- -

-

- -

-

3

g j j b j D b

D b

g j j b j D b

D b

g j j b j D b

D b

Dd

d

bj

d

1

85 4 4 2

,

1

42 6 4 2 4

,

1

82 7 4 4 2

27

8,

,

1 cos

1 2 cos.

jj j

j

jj j

j

jj j

jj

k kk

k

j

,452

1 2 1 2

21 2

,492

1 21

1 21

21 2

1

,532

1 22

1 22

21 2

2,3

1 20

2

2

( )

( ) (( ) ( ) ( ) ( )

( ))

( ) (( ) ( ) ( ) ( )

( ))

( ) (( ) ( ) ( ) ( )

( ))

( ) ( )

In the definitions of the gj xx, functions, we have neglected allindirect contributions which arise only at higher orders ofeccentricity (Murray & Dermott 1999). Here ai is thesemimajor axis, ei the orbital eccentricity, mi the mass, li the

mean longitude, and vi the longitude of periastron of the -i thplanet (i= 1 or 2), a = a a1 2 and M is the mass of the star.Throughout this derivation, we will make use of the

following small quantities:

dw

=

=

nm

me

,

,

, 4

j

i

ii

i ( )

where

=L

nG M m

5ii

i

2 2 3

3( )

and

w = - -jn j n2 . 6j 2 1( ) ( )

The assumption that the masses are small is required becausewe neglect terms of order i

2 in both Equation (1) and below,for example in Equations (9)–(12). The assumption of loweccentricities allows us to neglect higher-order terms inEquation (1), as well as terms at zeroth and linear ineccentricities that are non-resonant. We will further discussthe assumptions of near resonance, d 1, and of loweccentricities, e 1i , below. For reference, δ is on the orderof a few percent for many systems of interest.We will derive the TTVs using an approach developed by

Nesvorný & Morbidelli (2008), Nesvorný (2009), andNesvorný & Beaugé (2010) and later used by Deck & Agol(2015). The method is based on perturbation theory within aHamiltonian framework; therefore, we need to first convert theorbital elements into canonical variables. Written in terms ofthe canonical momenta (left) and coordinates (right),

lL =

= =

m GM a

x P p y P p2 cos 2 sin

i i i i

i i i i i i

with

v= L + = -P O e p ,i ie

i i i24i

2

( )

Equation (1) can be rewritten as

q q

= L L +

=-L

-L

=-L

+

=L

- +L

-

+LL

-

=-L

-L

-LL

+

H H H

HG M m G M m

HG M m

A A

Ag

x yg

x y

gx x y y

Ag

x yg

x yg

x y x y

, ,

2 2,

cos sin ,

,

2 2 .

7

j j

j j

j

j j j

0 1 2 1

0

2 213

12

2 223

22

1 1

2 223

22 1 2

1,45

112

12 ,53

222

22

,49

1 21 2 1 2

2,45

11 1

,53

22 2

,49

1 21 2 2 1

( )

[ ˜ ˜ ]

˜ ( ) ( )

( )

˜ ( )

( )

The coefficients gj xx, are evaluated at semimajor axis

ratio a L L = L Lm m,1 2 1 12

2 22( ) ( ) ( ) .

2

The Astrophysical Journal, 821:96 (13pp), 2016 April 20 Deck & Agol

Page 3: TRANSIT TIMING VARIATIONS FOR PLANETS NEAR

Our goal is to find a new set of canonical variables (denotedwith primes) such that in the new set the Hamiltonian takes theform

¢ = L¢ L¢ = -L

-L¢ ¢

H HG M m G M m

,2 2

. 80 1 2

2 213

12

2 223

22

( ) ( )

In the new variables, the motion of the two planets is“Keplerian.” These Keplerian orbits correspond to theaverage of the perturbed orbits (averaged over the periodicterms in H1). TTVs are deviations from the times predictedfrom the mean ephemeris of a planet. In practice we estimatethis by fitting the transit times with a constant period(Keplerian) model. The deviations are caused by the interactionwith the other planet, and hence the transformation we seek toturn Equation (1) into (7) is precisely what we need to give usthe TTVs.

To determine this transformation, we use a Type-2generating function of the form

l l lL¢ ¢ = L¢ + ¢ + L¢ ¢F y x y x f y x, , , , , , , 9i i i i i i i i i i i i2 ( ) ( ) ( )

which relates the old and new variables as

l l

l l

L =¶¶

= L¢ +¶¶

=¶¶

= ¢ +¶¶

¢ =¶¶L¢

= +¶¶L¢

¢ =¶¶ ¢

= +¶¶ ¢

F f

xF

yx

f

y

F f

yF

xy

f

x

,

,

,

. 10

ii

ii

ii

ii

ii

ii

ii

ii

2

2

2

2 ( )

Therefore, the first piece of F2 is the identity transformationand the second piece, f, is a small correction which will belinear in 1. The function f which produces the new “Keplerian”Hamiltonian of Equation (8) from the Hamiltonian ofEquation (7) is

w

q q=L

-fn

A Asin cos , 11j

j j12 2

1 2[ ˜ ˜ ] ( )

which we determine by solving the homologic equation

+ =f H H, 0 120 1{ } ( )

where ¼ ¼,{ } denotes a Poisson Bracket. For more details onthis type of derivation, one can refer to Morbidelli (2002),Nesvorný & Morbidelli (2008) or Deck & Agol (2015).Formally, f is a mixed-variable function of both old coordinatesand new momenta. In Equation (11), however, we neglect tomake this distinction. This is because the difference betweenthe two sets, given implicitly in Equation (10), depends onderivatives of f, which is itself linear in 1. Therefore, within fitself, the difference between the two sets is negligible since weare only working to first order in 1.

Equation (10) shows us how to derive, using the function f ofEquation (11), the difference between the real and averaged

canonical variables. The results are

⎡⎣⎢

⎤⎦⎥

⎡⎣⎢

⎤⎦⎥

l l dlw

wq q

dw

q q

dw

q q

d

- ¢ º »L

L-

- ¢ º = -L

-

- ¢ º =L

-

L - L¢ º L »

n d

dA A

y y yn dA

dx

dA

dx

x x xn dA

dy

dA

dy

sin cos ,

sin cos ,

sin cos ,

0. 13

i i ij

j

ij j

i i ij i

ji

j

i i ij i

ji

j

i i i

12 2

2 1 2

12 2 1 2

12 2 1 2

[ ˜ ˜ ]

˜ ˜

˜ ˜

( )

Here we have made use of the small parameters given inEquation (4). That is, we have assumed that since the pair isclose to resonance with small eccentricities, we need onlyretain terms of the order de and de 2( ) . Terms proportional to

de2 or missing the small denominator δ are assumed to beconsiderably smaller and can be neglected. This approximationholds as long as e 1 and as long as d 1. The dominantneglected term de2 is negligible compared with the synodicchopping terms (of zeroth order in e and d1 ) for de . Forδ of a few percent, as for many of the Kepler systems, thiscorresponds to e 0.1.We now must take changes in the canonical elements and

convert them into TTVs. The transit occurs when the trueanomaly of the planet θ is equal to a value, which, given thereference frame, aligns the planet in front of the star along ourline of sight. θ is not a canonical variable, but it can be relatedto our canonical set via a power series in eccentricity of thetransiting planet:

q l l l lL = +L

+ + + ¼x y x y O e, , ,2

sin cos

14

2[ ] ( ) ( )

( )

Perturbing Equation (14) about the averaged orbit yields

dqq l

ldl

q ld

q ld

=¶ L

¶+

¶ L¶

+¶ L

x y x y

xx

x y

yy

, , , , , ,

, , ,15

[ ] [ ]

[ ] ( )

where we have already neglected the dL piece with theknowledge that it will be small. The derivatives ofq l L x y, , ,[ ] with respect to any of the remaining variableswill be proportional to e e e, ,0 1 2without any small denomi-nators. Hence, we also only keep q l¶ L ¶x y x, , ,[ ] ,q l¶ L ¶x y y, , ,[ ] , and q l l¶ L ¶x y, , ,[ ] to zeroth order in

e since we have assumed eccentricities are small. Weapproximate

dq dl d l d l» +L

+ + ¼x y2

sin cos . 16( ) ( )

To turn Equation (16) into a timing perturbation, we need toconvert dq into dt. This is achieved by relating θ to λ, since λ isa linear function of time. We can write

dq d= - + +¼n t O e , 17( ) ( )

where again we can neglect the O(e) correction, since this is afactor of e without a small denominator d w= nj i. Combiningthe results of Equations (13), (16), and (17) we find that the

3

The Astrophysical Journal, 821:96 (13pp), 2016 April 20 Deck & Agol

Page 4: TRANSIT TIMING VARIATIONS FOR PLANETS NEAR

TTVs are approximately given by

⎣⎢⎢

⎧⎨⎩

⎛⎝⎜

⎞⎠⎟

⎛⎝⎜

⎞⎠⎟

⎫⎬⎭

⎧⎨⎩

⎛⎝⎜

⎞⎠⎟

⎛⎝⎜

⎞⎠⎟

⎫⎬⎭

⎦⎥⎥

d aw w

q

w wq

=- - -

+ - -

tn

jn

An

B

jn

An

B

13 2 2 sin

3 2 2 cos 18

j jj

j jj

11

21

2

11

11

12

21

21

( )

( ) ( )

and

⎣⎢⎢

⎧⎨⎩

⎛⎝⎜

⎞⎠⎟

⎛⎝⎜

⎞⎠⎟

⎫⎬⎭

⎧⎨⎩

⎛⎝⎜

⎞⎠⎟

⎛⎝⎜

⎞⎠⎟

⎫⎬⎭

⎦⎥⎥

dw w

q

w wq

=- - -

+ - -

tn

jn

An

B

jn

An

B

13 2 sin

3 2 cos 19

j jj

j jj

22

12

2

12

12

22

22

22 ( )

where

a v a v

a v v

a v a v

a v v

a l v a l v

a l v a l v

a l v a l v

a l v

a l v

= = +

+ +

=- = - -

- +

= + + +

=- + - +

= + + +

=- +

- +

A A g e g e

g e e

A A g e g e

g e e

B g e g e

B g e g e

B g e g e

B g e

g e

cos 2 cos 2

cos

sin 2 sin 2

sin

2 cos cos

2 sin sin

2 cos cos

2 sin

sin .

20

j j

j

j j

j

j j

j j

j j

j

j

1 1 ,45 12

1 ,53 22

2

,49 1 2 1 2

2 2 ,45 12

1 ,53 22

2

,49 1 2 1 2

11

,45 1 1 1 ,49 2 1 2

21

,45 1 1 1 ,49 2 1 2

12

,53 2 2 2 ,49 1 2 1

22

,53 2 2 2

,49 1 2 1

˜ ( ) ( ) ( ) ( )( ) ( )

˜ ( ) ( ) ( ) ( )( ) ( )

( ) ( ) ( ) ( )

( ) ( ) ( ) ( )

( ) ( ) ( ) ( )

( ) ( )( ) ( )

( )

When will these formulae break down? Here we give somequalitative expectations before turning to numerical tests inSection 4.

Although we are considering the near resonance case, if δbecomes too small neglected terms proportional to d1 willbecome important if the system is too close to resonance.Additionally, our formulae will not apply when a system is inresonance. The width of the resonance grows with e and with ò.For a given ò, then, we expect our formulae to fail if theeccentricity is too large, either because of being in resonance orbecause of neglected higher-order terms.

As a 1, the functions of Laplace coefficients appearing ascoefficients in the disturbing function can diverge, mitigatingthe effect of small eccentricities raised to high powers. Inpractice, this means that the derived formulae incur larger erroras a 1 due to these neglected higher-order in e terms.

Throughout this derivation, we neglected terms without asmall denominator δ. In the low eccentricity regime, thedominant contribution comes from the terms independent ofeccentricity. If de , these zeroth order (in e) chopping termswill be comparable in magnitude to the second-order TTV. Inpractice, this implies that for many systems the second-orderresonant formula should be used with the chopping formula ofNesvorný & Vokrouhlický (2014) or Deck & Agol (2015). Ofcourse, if the second-order MMR under consideration is aO e2( ) correction of a first-order MMR, one must include thefirst-order resonant contributions (not presented here) as well.

Note that these TTVs were derived in terms of Jacobielements, which, for the outer planet are not the defined relativeto the star but rather to the center of mass of the inner-planet-star system. However, true transits occur with respect to thestar, not the center of mass of the inner subsystem. Thenecessary correction can be determined by treating the motionof the star as a sum of two Keplerian orbits (e.g., Agolet al. 2005). However, the indirect contributions resulting fromthis correction do not have small denominators, and hence theyare not important at this level of approximation.

3. INTERPRETATION OF THE APPROXIMATE TTV

The TTV expressions given in Equations (18) and (19)depend on the eccentricities and longitudes of pericenter ofeach planet after averaging over the TTV period and the orbitalperiods of the planets.4 The only further variation in thesequantities is due to secular evolution. If we assume that theobservational baseline is short compared with the seculartimescale, they will be approximately constant. We now showthat the TTVs approximately depend only on the masses of thetwo planets (relative to the mass of the host star) and theapproximately constant quantities

dd

= -= -

k k kh h h 21

1 2

1 2 ( )

where v=h e sini i i and v=k e cosi i i.We define

wD º

-- = -

j

j

P

P jn

21 , 22j

j2

1 2

( )

and, substituting in, the TTVs become

⎡⎣⎢⎢

⎧⎨⎩⎫⎬⎭

⎧⎨⎩⎫⎬⎭

⎤⎦⎥⎥

da

a q

a q

=-D

-D

+

+-D

+

-

-

tn j

j

jA B

j

jA B

2 3 2

2sin

3 2

2cos 23

j jj

jj

11

1 2 23 2

1 11

3 22 2

1

( )

( ) ( )

and

⎡⎣⎢⎢

⎧⎨⎩⎫⎬⎭

⎧⎨⎩⎫⎬⎭

⎤⎦⎥⎥

d q

q

=-D

-D

+

+ -D

+

tjn

A B

A B

2 3

2sin

3

2cos . 24

j jj

jj

22

1 1 12

2 22 ( )

Next, we choose our reference frame such that the truelongitude θ at transit is zero.5 Then in accordance with ourprevious neglect of terms of O(e) without a small denominatorδ, the mean longitude at transit is also zero.Now, as shown by Figure 1, with an error of only a factor of

1–2 one can approximate a a»g gj j,53 ,45( ) ( ) anda a» -g g2j j,49 ,45( ) ( ) (shown evaluated at

4 To be clear, the “average” eccentricity referred to here and going forward isthe eccentricity computed from the average canonical variables x and y, whichdiffers from the average in time of the eccentricity.5 Note that we are working in Jacobi coordinates and that in reality the transitoccurs when the true longitude in astrocentric coordinates is zero. However,here we make an approximation to λ at transit at the TTV level, which isalready of order ò. The correction from Jacobi to astrocentric coordinates for λat transit, of order ò, therefore produces an 2 correction that we can safelyignore.

4

The Astrophysical Journal, 821:96 (13pp), 2016 April 20 Deck & Agol

Page 5: TRANSIT TIMING VARIATIONS FOR PLANETS NEAR

a = -j j2 2 3[( ) ] ), in which case

a d d

a d d

a d

a d

a d

a d

» -

»-

»

»-

»- = -

» = -

A g k h

A g k h

B g k

B g h

B g k B

B g h B

2

2

2

2

2 . 25

j

j

j

j

j

j

1 ,452 2

2 ,45

11

,45

21

,45

12

,45 11

22

,45 21

( )[ ]( )

( )

( )

( )

( ) ( )

Then the TTVs are roughly given by

⎡⎣⎢⎢

⎧⎨⎩⎫⎬⎭

⎧⎨⎩⎫⎬⎭

⎤⎦⎥⎥

da

d d d q

d d d q

=-D D

- +

-D

+

tg

n jk h k

k h h

2 3

22 sin

32 cos

26

j

j jj

jj

1,45

11 2 2

2 2[ ]

( )

and

⎡⎣⎢⎢

⎧⎨⎩⎫⎬⎭

⎧⎨⎩⎫⎬⎭

⎤⎦⎥⎥

d d d d q

d d d q

=-D

-D

- +

+D

+

tg

jnk h k

k h h

2 3

22 sin

32 cos , 27

j

j jj

jj

2,45

21

2 2[ ]

( )

where we have also approximated a » -- j j 23 2 ( ). If we set

⎧⎨⎩⎫⎬⎭

⎧⎨⎩⎫⎬⎭

f

d d d

f

d d d

=D

- +

=D

+

k h k

k h h

cos

3

22

,

sin

32

,

j

j

2 2[ ]

and

⎧⎨⎩⎫⎬⎭

⎧⎨⎩⎫⎬⎭ d d d d d d=

D- + +

D+k h k k h h

3

22

32 ,

28j j

2 2 2

2 2

[ ]

( )

then

da

pq f= -

D-

-

tP g

jsin 29

j

jj1

1 ,451 2

2 ( ) ( )

and

dp

q f=D

-tP g

jsin . 30

j

jj2

2 ,451 ( ) ( )

Therefore, the TTVs of a pair of planets near a second-orderresonance with low eccentricity are approximately given bysinusoidal motion with a phase f set by the eccentricities andpericenters in the combinations of d v v= -k e ecos cos1 1 2 2

and d v v= -h e esin sin1 1 2 2. The amplitude is determinedby both the mass of the perturbing planet and a factordepending again on dk and dh. Finally, the TTVs of the twoplanets are anti-correlated. Assuming both planets transit(a D P P, , ,j 1 2 and the time evolution of qj are known), theonly unknowns are , ,1 2 , and f. However, from the TTVsalone, we only obtain two amplitudes and a phase.This outcome is similar to that of the TTVs of a pair of

planets near first-order resonances, where only the combina-tions dh and dk appeared in the TTVs (through the real andimaginary parts of = +v vZ fe e ge ei i

free 1 21 2, with f≈−g).

The relation f≈−g is analogous to the approximation wemade here regarding Laplace coefficients and the combinations

»g gj j,45 ,53 and » -g g2j j,49 ,45. Again, in that case theobservables are two amplitudes and a single phase becausethe TTVs are approximately anti-correlated.One way to break degeneracies is to measure other

independent harmonics in the TTVs. For example, one mightmeasure the “chopping” signal and determine 1 and/or 2 fromthat. In reality, there will be a nonzero phase offset, and if it ismeasured significantly, the TTVs yield four observables,assuming the mean ephemerides are known because eachplanet transits. However, this is still not enough to determineboth eccentricities and longitudes of pericenter uniquely. Itcould be that chopping effects which appear at first-order in theeccentricities allows individual eccentricities and longitudes tobe measured, though these are small amplitude.For a pair of planets near a first-order resonance, the TTV

derived above will represent an O(e2) correction to the TTVs.For example, a pair near the k: -k 1 resonance will exhibitTTVs with a period equal to p - -kn k n2 12 1∣ ( ) ∣ and also atthe second harmonic, with a period of p - -jn j n2 22 1∣ ( ) ∣(with =j k2 ). This second harmonic appears with a differentdependence on dh and dk , and if measured could allow forunique mass measurements as well. However, since both thefirst- and second-order harmonics depend approximately on dhand dk , the higher-order eccentricity corrections to the formulaof Lithwick et al. (2012) may not allow for individualeccentricity measurements, especially for low signal-to-noisedata. However, if the relative phase offsets of either harmonicfrom π can be measured, this second-order harmonic could, in

Figure 1. Validity of the approximation that a a»g gj j,53 ,45( ) ( ) andthat a a» -g g2 .j j,49 ,45( ) ( )

5

The Astrophysical Journal, 821:96 (13pp), 2016 April 20 Deck & Agol

Page 6: TRANSIT TIMING VARIATIONS FOR PLANETS NEAR

theory, allow for unique measurements of both eccentricitiesand pericenters as well. We test this in Section 4.2.3.

4. NUMERICAL TESTS OF THE FORMULA

4.1. Comparisons with Direct n-body Integration

We have carried out a comparison of the second-orderformula with n-body simulations of TTVs using TTVFast(Deck et al. 2014). We simulated a system of two planets with

= = -m m m m 101 25 with aligned longitudes of pericenter,

v v=1 2, and anti-aligned longitudes of pericenter,v v p= +1 2 . The initial phases and v1 were chosenrandomly. The TTVs determined from the n-body simulationwere computed for an inner planet period of 30 days, over aduration of 1600 days, to mimic a typical transiting planetsystem in the Kepler data set. The eccentricity vectors wereheld fixed at the value computed from the n-body simulationaveraged over 1600 days. The ephemerides were allowed tovary, and were varied to optimize the agreement between then-body and analytic formula, while α used in computing thecoefficients was given by a = P P1 2

2 3( ) , where P1 and P2 arethe periods fit to give the ephemerides.

We first focused on a range of aD( ) near the 5:3 second-order resonance. In Figure 2 we show, for anti-alignedpericenters, the fractional error in the formula given inEquation (18) and in (19). The error is less than 10% onlyfor a small range in eccentricity, but importantly the regionwhere the second-order formula alone applies is qualitatively asexpected. For low eccentricities, the chopping terms are aslarge as the second-order resonant terms and they are neglectedin this fit (as discussed at the end of Section 2). For largereccentricities, this narrow range of Δ includes resonant orbits,to which our formulae do not apply.

We also tried fitting the numerically determined TTVs withthe second-order formulae added to the first-order formulaepresented in Agol & Deck (2015). Note that the first-orderformulae include all terms linear in the eccentricity, while thesecond-order formulae only include the near-resonant pieces. InFigure 3 we show the resulting comparison between theextended formula and the n-body results. The agreement is nowexcellent even at low eccentricity, as expected since we haveincluded the chopping terms. However, there is still a clearresonant region where our formulae fails.Figures 4 and 5 show the results of the aligned and anti-

aligned longitudes of periastron, now for a much larger rangeof α. The anti-aligned longitudes of periastron tends tomaximize the discrepancy, while the aligned longitudes tendto minimize it. The fractional precision of the model wascomputed from the scatter of the residuals of the fit divided intothe scatter in the n-body TTVs. The mean longitudes and thelongitude of periastron of the inner planet were chosenrandomly and do not affect the appearance of this plotsignificantly.We found that including the second-order term improves the

fit to the n-body simulation significantly near j:j−2 periodratios (as demonstrated also by Figures 2 and 3), which areindicated in the plot, allowing the analytic formulae to be usedto much higher eccentricity than in the case of the first-orderformula only (Agol & Deck 2015). It also improves the fit nearthe j:j−1 resonances as each of these is close to a 2j:2j−2resonance, and thus can be affected by the second-order ineccentricity terms.

4.2. Applications to Real and Simulated Systems

We next explored how parameter estimates of masses,eccentricities, and longitudes of periastron derived by fittingboth real and simulated data with our formulae compared to

Figure 2. Comparison of TTVFast with the second-order formula near the 5:3second-order resonance. Error is given by the standard deviation of theresiduals of the analytic fit to the numerical TTVs, divided by the standarddeviation of the TTVs. The top panel shows the result for the inner planet, thebottom is for the outer planet. The dotted lines indicate the 10% error level.

Figure 3. Comparison of TTVFast with the first-order plus second-orderformula near the 5:3 second-order resonance. Error is given by the standarddeviation of the residuals of the analytic fit to the numerical TTVs, divided bythe standard deviation of the TTVs. The dotted lines indicate 10% error level.

6

The Astrophysical Journal, 821:96 (13pp), 2016 April 20 Deck & Agol

Page 7: TRANSIT TIMING VARIATIONS FOR PLANETS NEAR

those found using TTVFast. We used the second-order resonantterms, in combination with the first-order eccentric formulae ofAgol & Deck (2015), as our analytic model unless otherwisenoted. For one simulated system we also used the first-orderformulae alone for a test. In the following analyses, weemployed an affine invariant Markov chain Monte Carlo(Goodman &Weare 2010) to estimate parameters. When fittingreal data, we used a student-t likelihood function with 2 degreesof freedom and did not remove outliers. When fitting simulatedtransit times with Gaussian uncertainties added, we used aGaussian likelihood function.

4.2.1. Kepler-26 (KOI-250)

The Kepler-26 star hosts four planets with orbital periods of3.54 (Kepler-26d), 12.28 (Kepler-26b), 17.26 (Kepler-26c),and 46.8 (Kepler-26e) days. The period ratios of adjacentplanets are 3.47 (d–b), 1.41 (b–c), and 2.71 (c–e). Theinnermost planet and outermost planet, therefore, are not near alow-order MMR with either middle planet. They do not exhibitTTVs of their own and we assume that they do not affect theTTVs of the middle two planets either. In that case, the b–c paircan be treated as an isolated system near the 7:5 resonance. TheTTVs of these two planets show periodic behavior on atimescale given by the “super period” p -n n2 7 52 1∣ ∣ asexpected, along with a synodic TTV signal (Jontof-Hutteret al. 2016).We modeled the b–c pair using an analytic model that

included the 7:5 second-order MMR terms and all terms thatappear at first and zeroth order in eccentricity. We obtainplanet-star mass ratio measurements of =Å M M M Mb( )( )

Figure 4. Comparison of TTVFast with the first-order plus second-orderformula. Error is given by the standard deviation of the residuals of the analyticfit to the numerical TTVs, divided by the standard deviation of the TTVs withthe longitudes of periastron aligned (v v=1 2). The top panel shows the resultfor the inner planet, the bottom for the outer planet. Dotted lines indicate 10%error level. Upper right: Hill unstable models were not computed, and show100% error. Green dashed lines indicate locations of j:j−1 resonances; bluedashed lines show j:j−2 resonances.

Figure 5. Comparison of TTVFast with the first-order plus second-orderformula. Error is given by the standard deviation of the residuals of the analyticfit to the numerical TTVs, divided by the standard deviation of the TTVs withthe longitudes of periastron anti-aligned (v v p= +1 2 ). The labels and linesare the same as in Figure 4.

Figure 6. Joint posterior probability distribution for the masses of Kepler-26band Kepler-26c, in Earth masses, assuming a solar mass star. The best-fit valuesfrom the formula fit are denoted with the red point, with 68% (red) and 95%(black) confidence contours shown as well. The blue point and dashed linesreflect the best fit and 68% boundaries of the Jontof-Hutter et al. (2016) fitusing a student-t likelihood function.

7

The Astrophysical Journal, 821:96 (13pp), 2016 April 20 Deck & Agol

Page 8: TRANSIT TIMING VARIATIONS FOR PLANETS NEAR

9.78 1.36 and = Å M M M M 11.92 1.38c( )( ) , whichare in close agreement with those obtained in a full dynamicalanalysis of the data (Jontof-Hutter et al., under review), asshown in Figure 6.

Using the formulae, we find two possible linear correlationsbetween k1 and k2 (where v=k e cosi i i) and between h1 and h2(where v=h e sini i i). These arise since the second-order TTVdepends approximately on quadratic functions of d = -k k k1 2and d = -h h h1 2. That is, dk2 and dh2 are singly peaked, andhence dk and dh can be either positive or negative, leading totwo linear correlations between k1 and k2 (and between h1 andh2). The slope we find that fits these correlations is near unity,as expected based on the arguments in Section 3, and the smalldeviation is related to the error incurred by approximating

» -g g2j j,49 ,45 and =g gj j,53 ,49. One of these modes ispreferred compared with the other, likely because of (singlemode) constraints on dk and dh resulting from the nearby 4:3and 3:2 resonances.

4.2.2. Kepler-46 (KOI-872)

The star Kepler-46 hosts two transiting planets with orbitalperiods of 6.8 (Kepler-46d) and 33.6 (Kepler-46b) days.Additionally, Kepler-46b exhibits TTVs due to a non-transitingcompanion (Kepler-46c) near the 5:3 resonance (Nesvornýet al. 2012). We modeled the transit times of Kepler-46bpresented in Nesvorný et al. (2012) using the second-orderterms for the 5:3 resonance in combination with the full first-order formula of Agol & Deck (2015) and ignoring Kepler-46 d. We held the mass of Kepler-46b fixed, as its own massdoes not affect its TTVs, but the mass of Kepler-46c and theeccentricities, arguments of pericenter, periods, and orbitalphases of each planet were allowed to vary. We only searchedfor a solution near the 5:3 resonance with Kepler-46b,however.

In Figure 7 we show the results of our formula fit to thetransit times of Kepler-46b in comparison with numericalresults determined by Nesvorný et al. (2012). We find closeagreement in these parameters, as well as in the upper limit of∼0.02 found for the eccentricity of Kepler-46b (not shown).Given the best-fit period ratio of 1.70, the expected superperiod for the 5:3 MMR is nearly 18 orbits of Kepler-46b. Thedata extend a baseline of 15 transits, with periodicity of ∼5–6orbits of Kepler-46b and no clear large amplitude signal at aperiod of 18 orbits. The constraints in this system, therefore,likely come entirely from a strong “chopping” TTV, enhancedby contributions from the distant 3:2 MMR and 2:1 MMR(with contributions at periodicities of several orbital periods, asnoted by Nesvorný et al. 2012). The constraints on theeccentricities likely come from the magnitude of the 3:2 and2:1 MMR contributions as well as the lack of a large signal dueto the 5:3 resonance.

4.2.3. Fits to Simulated Data for a System Near a First-order MMRand Prospects for Measuring Individual Eccentricities

Our test case consisted of two 5 Earth-mass planets orbitinga solar mass star, with initial osculating periods of 10.0 and20.2 days, eccentricities of 0.035 and 0.05, and longitudes ofpericenter misaligned by 135°. For these parameters, thesystem is near the 2:1 resonance, but with importantcontributions from the 4:2 resonance due to the moderateeccentricities and values of dk and dh. We remark that this

system is somewhat similar to KOI-142, which has two planetsnear a 2:1 resonance with eccentricities of ∼0.05, and whichalso exhibits TTVs that deviate from a pure sinusoid (Nesvornýet al. 2013).We simulated transit times using TTVFast and added

Gaussian noise with a standard deviation of two minutes. InFigure 8 we show the modeled TTVs. The various coloredpoints show a sample fit found modeling this data with thesecond-order resonant terms and the first-order resonant terms.We also show the contribution of this fit coming from the first-order resonant terms alone, as well as from the second-orderresonant terms alone (note that the second-order formulaecontains the O(e) contribution of the second-order resonantterms; we do not double count this).In Figures 9 and 10 we show the resulting joint confidence

levels for the two planet masses and eccentricities for both theformula model and for an n-body model. The agreement is verygood; we note, however, that modes associated with highereccentricity were also found using the formula fit, dependingon the particular noise realization. However, this multi-modality disappeared as the noise amplitude decreased. Notealso that we used direct n-body integration to simulate the

Figure 7. Posterior probability distribution for the perturber (Kepler-46c) massand eccentricity determined using the formula (black). The blue solid lineindicates the best fit of Nesvorný et al. (2012) and the dashed lines reflect the±34% confidence contours.

8

The Astrophysical Journal, 821:96 (13pp), 2016 April 20 Deck & Agol

Page 9: TRANSIT TIMING VARIATIONS FOR PLANETS NEAR

transit times we fit. Hence the input value for the eccentricity isan osculating value. The eccentricities measured via theformula correspond to eccentricities computed from theaveraged (canonical) x and y variables, or free eccentricities,while those of the n-body model are initial osculating elements.The difference between the two is on the order of themagnitude of the forced eccentricity due to the nearbyresonance, and may explain why there is a small offsetbetween the numerical and n-body fits.

For these parameters and signal-to-noise, the eccentricities(and longitudes of pericenter, not shown) are both measuredindependently. As discussed in Section 3, the first-orderresonant terms as presented in Lithwick et al. (2012) alone

suffer from an absolute degeneracy, in that the TTV amplitudeand phase depends only on quantities approximately equal tod v v= -h e ecos cos1 1 2 2 and d v v= -k e esin sin1 1 2 2 (thereal and imaginary parts of Zfree in the notation of Lithwicket al. 2012). Though both the first- and second-order TTVharmonics depend approximately only on dh and dk , in reality,they both depend on slightly different functions of theeccentricity and pericenter.6 Hence at high signal-to-noise thetwo harmonics may be used to measure eccentricities andpericenters individually, as in Figure 10. Similarly, wehypothesize that the moderate eccentricities of the KOI-142system produce a detectable second harmonic in the TTVs ofKOI-142b, which, in combination with the short periodchopping and the transit duration variations, help lead to aunique solution for the non-transiting perturber (Nesvornýet al. 2013).Note that a model neglecting the second harmonic fails to

determine the eccentricities correctly: it returns a decent fit, butat significantly higher eccentricities. Including the second-orderterms in the model leads to the correct answer, but asmentioned there can be multi-modality. It is possible thatincluding the 6:3 resonant harmonic in the fit (an O e3( )correction, see below) would alleviate this.To explore how the second-order harmonic can lead to an

eccentricity measurement, as in Figure 10, we decreased theeccentricities to 0.014 and 0.01 in order to reduce the amplitudeof the second-order harmonic. All other parameters remainedthe same. In this case, a 2 minute amplitude for the noise is2–4x larger than the amplitude of the second-order harmonicfor the two planets. Figure 11 shows the transit times wemodeled in addition to a sample fit, again delineating betweenthe entire second-order model and the contributions comingfrom the first-order resonant piece and the second-order piece.Figures 12 and 13 show the result of a fit to these times using

n-body, using the second- and first-order eccentricity terms,

Figure 8. Simulated transit timing variations for the higher eccentricity case(black), a representative solution from the model including the second-orderand first-order terms (from Agol & Deck 2015; orange), the contribution to thismodel from the first-order resonant terms alone (turquoise triangles), and thecontribution from the second-order resonant harmonic (in red). Not shown isthe individual contribution coming from “chopping” terms without any smallresonant denominators.

Figure 9. Joint posterior 68% (dotted) and 95% (solid) confidence contours forthe planet masses, in units of Earth mass. The results shown are from a fulldynamical analysis (blue) and from an analysis using the second-order terms incombination with the first-order formulae derived in Agol & Deck (2015;black).

Figure 10. Joint posterior 68% (dotted) and 95% (solid) confidence contoursfor the planet eccentricities. The results shown are from a full dynamicalanalysis (blue) and from an analysis using the second-order terms incombination with the first-order formulae derived in Agol & Deck (2015;black).

6 This is true especially for the 2:1 MMR, which contains indirect terms in thecoefficients of the first-order resonant terms, as noted by Hadden &Lithwick (2015).

9

The Astrophysical Journal, 821:96 (13pp), 2016 April 20 Deck & Agol

Page 10: TRANSIT TIMING VARIATIONS FOR PLANETS NEAR

and now also with only the first-order eccentricity formula.Apparently the second-order terms are still a useful constrainton the eccentricities and longitudes of pericenters, even giventheir low amplitude. This may be because the amplitude of thesecond-order term is constrained to be below the noise. Forthese particular data, the planet masses are measured equallywell by all models, as shown Figure 14.

5. HIGHER-ORDER RESONANCES

We now extend our above derivation to any ordereccentricity-type resonance. This generalizes the work of Bouéet al. (2012), who studied the particular case of a planet on aninitially circular orbit and a planet on a fixed eccentric orbit. Inthis case, for the j:j−N resonance, the Hamiltonian, including

only the resonant terms, takes the form

å a q f

=- - -

´ -=

-

HGM m

a

GM m

a

Gm m

a

g e e

2 2

cos 31k

N

j k Nk N k

j N k

1

1

2

2

1 2

2

0, ; 1 2 ;( ) ( ) ( )

where

q l lf v v

= - -= + -

j j N

k N k

,

, 32j N

k

; 2 1

1 2

( )( ) ( )

and agj k N, ; ( ) is a function of Laplace coefficients. For example,for the 2nd order j:j−2 resonance, =k 0, 1, 2, and in thatcase = =g g g g,j j j j,0;2 ,53 ,1;2 ,49 and =g gj j,2;2 ,45. We rewrite the

Figure 11. Simulated transit timing variations for the lower eccentricity case(black), a representative solution from the model including the second-orderand first-order terms (from Agol & Deck 2015; orange), the contribution to thismodel from the first-order resonant terms alone (turquoise triangles), and thecontribution from the second-order resonant harmonic (red).

Figure 12. Joint posterior 68% (dotted) and 95% (solid) confidence contoursfor the planet eccentricity vector components I. The results shown are from afull dynamical analysis (blue) and from an analysis using the second-orderterms in combination with the first-order formulae derived in Agol & Deck(2015; black), and one using only the first-order solution (red).

Figure 13. Joint posterior 68% (dotted) and 95% (solid) confidence contoursfor the planet eccentricity vector components II. The results shown are from afull dynamical analysis (blue) and from an analysis using the second-orderterms in combination with the first-order formulae derived in Agol & Deck(2015; black), and one using only the first-order solution (red).

Figure 14. Joint posterior 68% (dotted) and 95% (solid) confidence contoursfor the planet masses, in units of Earth masses. See Figure 12 for details.

10

The Astrophysical Journal, 821:96 (13pp), 2016 April 20 Deck & Agol

Page 11: TRANSIT TIMING VARIATIONS FOR PLANETS NEAR

interaction Hamiltonian as

åa

q q

=- LL L

´ +f f=

--

R I

H ng

P P

e e

2 2

cos sin 33k

Nj k N

k N kk N k

ij N

ij N

1 1 2 20

, ;

12

22 1

22

2

; ;k k

( )( ) ( )

[ ( ) ( ) ] ( )

( )( )

where i not in a subscript position is -1 .We can simplify this:

åa

q q

=- LL L

´ -=

-

- -R I

H ng

z z z zcos sin 34

k

Nj k N

k N k

k N kj N

k N kj N

1 1 2 20

, ;

12

22

1 2 ; 1 2 ;

( )

[ ( ) ( ) ] ( )

( )

where

= + =z x iy P e2 35i i i iipi ( )

and we have remembered that the canonical angle is not ϖ butv= -p . This is the exact form of Equation (7), with

å

å

a

a

=L L

=-L L

=-

-

=-

-

R

I

Ag

z z

Ag

z z . 36

k

Nj k N

k N kk N k

k

Nj k N

k N kk N k

10

, ;

12

22 1 2

20

, ;

12

22 1 2

˜ ( )( )

˜ ( )( ) ( )

( )

( )

If we set N=2, we find agreement with the expressions for A1̃

and A2˜ given in Equation (7). We now proceed exactly asabove, to find

⎡⎣⎢

⎤⎦⎥

⎡⎣⎢

⎤⎦⎥

dlw

wq q

dw

q q

dw

q q

d

»L

L-

=-L

-

=L

-

L »

n d

dA A

yn dA

dx

dA

dx

xn dA

dy

dA

dy

sin cos

sin cos

sin cos

0 37

ij N

j

ij N j N

ij N i

j Ni

j N

ij N i

j Ni

j N

i

12 2

;2 1 ; 2 ;

12 2

;

1;

2;

12 2

;

1;

2;

[ ˜ ˜ ]

˜ ˜

˜ ˜

( )

with w = - -jn j N nj N; 2 1( ) . The deviation in the truelongitude is given by (16). We now sketch the derivation forthe inner planet:

⎡⎣⎢⎢

⎧⎨⎩⎛⎝⎜

⎞⎠⎟

⎫⎬⎭

⎧⎨⎩⎛⎝⎜

⎞⎠⎟

⎫⎬⎭

⎤⎦⎥⎥

dqw

a qw

l l

qw

l l

= ´-

+ L -

+ --

+ L -

n n j NA

dA

dy

dA

dx

n j NA

dA

dx

dA

dy

sin3

2 sin cos

cos3

2 cos sin . 38

j Nj N

j N

j Nj N

1 21

;;

1

;1

11

11

1

11

;1

;2

12

11

2

11

( ) ˜

˜ ˜

( ) ˜

˜ ˜( )

The derivative of the real (imaginary) part of a function is thereal (imaginary) part of the derivative of the function, i.e.,

= -R Iix x( ) ( ) and =I Rix x( ) ( ) for a complex number x, so

we can write

å

å

å

å

L =

L =-

L =-

L =-

f v

f v

f v

f v

=

- - - -

=

- - - -

=

- - - -

=

- - - -

R

I

I

R

dA

dxg ke e e

dA

dyg ke e e

dA

dxg ke e e

dA

dyg ke e e 39

k

N

j k Nk N k i

k

N

j k Nk N k i

k

N

j k Nk N k i

k

N

j k Nk N k i

11

1 0, ; 1

12

11

1 0, ; 1

12

12

1 0, ; 1

12

12

1 0, ; 1

12

k

k

k

k

1

1

1

1

˜( )

˜( )

˜( )

˜( ) ( )

[ ]

[ ]

[ ]

[ ]

with analogous expressions for the outer planet (with the pre-factor -k N k , f v f v- -k k1 2, the exponent of e1becomes k, and that of e2 becomes - -N k 1.). Written interms of eccentricities and pericenters, A1̃ and A2˜ are

å

å

a

a

=

=-

f

f

=

- -

=

- -

R

I

A g e e e

A g e e e , 40

k

N

j k Nk N k i

k

N

j k Nk N k i

10

, ; 1 2

20

, ; 1 2

k

k

˜ ( ) ( )

˜ ( ) ( ) ( )

where f v v= + -k N kk 1 2( ) .The final expressions for the deviations in transit times are

(after some simplification)

⎡⎣⎢⎢

⎧⎨⎩⎫⎬⎭

⎧⎨⎩⎫⎬⎭

⎤⎦⎥⎥

⎡⎣⎢⎢

⎧⎨⎩⎫⎬⎭

⎧⎨⎩⎫⎬⎭

⎤⎦⎥⎥

å

å

dw

a a

wq

wq

dw

a

wq

w

q

= -

´-

-

+ --

+

= -

´ - - -

+ + -

´

f f v l

f f v l

f f v l

f f v l

=

- -

- +

- +

=

- -

- +

- +

R R

I I

R R

I I

41

tn

ng e e

n j Ne e k e

n j Ne e k e

tn

ng e e

jne e N k e

jne e N k e

32 sin

32 cos

32 sin

32

cos

j N k

N

j k NN k k

j N

i ij N

j N

i ij N

j N k

N

j k NN k k

j N

i ij N

j N

i i

j N

12

1

1

; 0, ; 2 1

1

1

;1 ;

1

;1 ;

21

2

2

; 0, ; 2

11

2

;2 ;

2

;2

;

k k

k k

k k

k k

1 1

1 1

2 2

2 2

( )

( )

( ) ( ) ( )

( ) ( ) ( )

( )

( ) ( ) ( )

( ) ( ) ( )

( )

( )

( )

( )

or, casting in the same symbols as we did for the second-orderresonances (Equation (18)),

⎣⎢⎢

⎧⎨⎩

⎛⎝⎜

⎞⎠⎟

⎛⎝⎜

⎞⎠⎟

⎫⎬⎭

⎧⎨⎩

⎛⎝⎜

⎞⎠⎟

⎛⎝⎜

⎞⎠⎟

⎫⎬⎭

⎦⎥⎥

d aw

wq

w wq

=- -

-

+ - -

tn

j Nn

A

nB

j Nn

An

B

13

2 sin

3 2 cos

42

j N

j Nj N

j N j Nj N

11

21

;

2

1

1

;11

;

1

;

2

21

;21

;

( )

( )

( )

11

The Astrophysical Journal, 821:96 (13pp), 2016 April 20 Deck & Agol

Page 12: TRANSIT TIMING VARIATIONS FOR PLANETS NEAR

and

⎣⎢⎢

⎧⎨⎩

⎛⎝⎜

⎞⎠⎟

⎛⎝⎜

⎞⎠⎟

⎫⎬⎭

⎧⎨⎩

⎛⎝⎜

⎞⎠⎟

⎛⎝⎜

⎞⎠⎟

⎫⎬⎭

⎦⎥⎥

dw w

q

w wq

=- - -

+ - -

tn

jn

An

B

jn

An

B

13 2 sin

3 2 cos 43

j N j Nj N

j N j Nj N

22

12

;

2

12

;12

;

2

;

2

22

;22

; ( )

where q l l= - -j j Nj N; 2 1( ) and

å

å

å

å

å

å

a f

a f

a f v l

a f v l

a f v l

a f v l

=

=-

= - +

=- - +

= - - +

=- - - +

=

-

=

-

=

- -

=

- -

=

- -

=

- -

A g e e

A g e e

B g ke e

B g ke e

B g N k e e

B g N k e e

cos

sin

cos

sin

cos

sin

44

k

N

j k NN k k

k

k

N

j k NN k k

k

k

N

j k NN k k

k

k

N

j k NN k k

k

k

N

j k NN k k

k

k

N

j k NN k k

k

10

, ; 2 1

20

, ; 2 1

11

0, ; 2 1

11 1

21

0, ; 2 1

11 1

12

0, ; 2

11 2 2

22

0, ; 2

11 2 2

( )

( )

( ) ( )

( ) ( )

( )( ) ( )

( )( ) ( )

( )

with f v v= + -k N kk 1 2( ) .Therefore, close enough to any eccentricity-type resonance

—with the caveat that the system is not in the resonance andthat the higher-order eccentricity terms neglected are small—the TTVs of a pair of planets are periodic with a timescale ofp - -jn j N n2 2 1∣ ( ) ∣. The amplitudes depend linearly on themass of the perturbing planet, relative to the mass of the star,and on the eccentricities and pericenters, as well as on the meanlongitude of the transiting planet at transit. The phases alsodepend on these quantities, though they are independent of themasses. However, the amplitude and phase of these TTVs donot uniquely constrain the masses, eccentricities, and pericen-ters, since this amounts to six parameters and only fourobservables.

We hypothesize that the TTVs will, in the limit of compactorbits (higher j for a given N), be anti-correlated, only dependon approximately dk and dh, and that the Nth order resonantTTV will include powers up to e N∣ ∣ , with d d=e k h,( ).Physically, this dependence makes sense since anti-alignedorbits allow for closer approaches and stronger interactionsbetween the planets. Mathematically this maximizes e∣ ∣ toproduce larger TTVs.

In the above derivation, we neglected terms of orderd-e eN 1( ) (where again δ is a normalized distance to

resonance, defined in Equation (4), and we assume it is small).More importantly, neglected chopping effects, without anysmall denominators, appear at every order in e. If one combinesthe Nth order resonant TTV with the e1 chopping TTVformulae (Agol & Deck 2015), one still will find errors at loweccentricity. This arises because e must be larger for higher-order resonances to be important, and in this case neglectedchopping terms at much lower powers of e may also beimportant. Hence the Nth order TTV formulae above may be oflimited use even if combined with other known formulae.

As an exercise, we can use these formulae to confirm thoseof Lithwick et al. (2012). In that case, N=1, and

v vv v

= + ==- - =

=

=

=

=

R

I

A ge fe Z

A ge fe Z

B f

B

B g

B

cos cos

sin sin

0

0 45

1 2 2 1 1 free

2 2 2 1 1 free

11

21

12

22

( )( )

( )

setting = +v vZ fe e ge ei ifree 1 2

1 2, = =f gj k N, 1; and = =g gj k N, 0;

and approximating λ(transit)=0. The TTVs then are

⎡⎣⎢

⎧⎨⎩⎫⎬⎭

⎧⎨⎩⎫⎬⎭

⎤⎦⎥

d a a

q a q

=-D

-D

+

´ +-D

- -

-

R

I

tn j

j

jZ f

j

jZ

2 1 3 1

2

sin3 1

2cos

46

j j

11

21 2 3 2

free

3 2free

( ) ( )

( ) ( )

( )

and

⎛⎝⎜

⎞⎠⎟

⎡⎣⎢

⎤⎦⎥

d

q q

=D D

-

´ +D

R

I

tn j

Z g

Z

2 1 3

2

sin3

2cos 47j j

22

1 free

free

{ }{ }

( )

( ) ( )

with D = - -j j P P1 12 1( ) ( ) .In the Lithwick et al. (2012) paper, the TTVs are written in

the form

d

q q

= +

= +

q

R I

tV

ie c c

V V2

. .

sin cos 48

i

j j

j

( ) ( ) ( )

Equations (46) and (47) take that form if

⎜ ⎟

⎛⎝⎜

⎞⎠⎟

⎛⎝

⎞⎠

pa

a

p

=-D

+D

-

=D

- +D

-VP

jf

j

jZ

VP

jg Z

3

2

1

3

249

11 2 1 2

3 2 free

22 1

free

( )

( )

which are equivalent to (A.28) and (A.29) of Lithwicket al. (2012).

6. CONCLUSION

We have derived an expression for the TTVs of a pair ofplanets near the j:j−2 second-order MMR in the regime oflow eccentricities. In this case, the TTV of each planet issinusoidal with a frequency of - -jn j n22 1( ) , an amplitudelinearly dependent on the mass of the perturbing planet, relativeto the mass of the star, and with both amplitude and phasedependent on a function of the eccentricities and longitudes ofpericenter. In this case, there are six parameters but only fourobservables, yielding (in principle) mass measurements but notunique eccentricity and pericenter measurements. We show thatthe same is true for higher-order eccentricity-type resonances.This second result, however, may not be of (much) practicaluse since few pairs are found very near high-order MMRswhere the formulae apply. However, it does illustrate that

12

The Astrophysical Journal, 821:96 (13pp), 2016 April 20 Deck & Agol

Page 13: TRANSIT TIMING VARIATIONS FOR PLANETS NEAR

TTVs of systems near an Nth order resonance will appear witha period given by the super period p - -jn j N n2 2 1∣ ( ) ∣ andtherefore that higher-order eccentricity corrections to the TTVsof planets near first-order resonances appear at harmonics ofthe fundamental (super) period = - -j P j P1 12 1∣ ( ) ∣.

At a further level of approximation, which will be relevantfor low signal-to-noise data, we have shown that the TTVs oftwo planets near a second-order resonance are anti-correlated.In this case, there is an explicit degeneracy between massesand the combinations v v-e ecos cos1 1 2 2 and v -e sin1 1

ve sin2 2. This result is entirely analogous to that found forfirst-order resonances. We hypothesize that this basic resultextends to higher-order eccentricity-type resonances.

To alleviate the degeneracies between parameters that resultfor near-resonant systems, one must measure a differentcomponent of the TTV. This could be the chopping signal,associated with each planet conjunction, which primarilydepends on the masses of the planets. The higher-ordercorrection associated with the 2k:2k−2 second-order resonancederived here can help constrain eccentricities of planets near thek: -k 1 first-order resonance, though high signal-to-noise datais likely required to lead to precise individual eccentricitymeasurements. When modeling TTVs, we find it helpful toconsider the number of significantly measured observables in aTTV (in terms of amplitudes, phases, etc. of differentharmonics) in comparison with the number of free parameters,in light of intrinsic degeneracies which TTV formulae help toilluminate.

We would like to thank the referee who helped us to clarifyand improve this document. K.D. acknowledges support from

the JCPA postdoctoral fellowship at Caltech. E.A. acknowl-edges support from NASA grants NNX13AF20G,NNX13AF62G, and NASA Astrobiology Institutes VirtualPlanetary Laboratory, supported by NASA under cooperativeagreement NNH05ZDA001C.

REFERENCES

Agol, E., & Deck, K. M. 2015, ApJ, 818, 117Agol, E., Steffen, J., Sari, R., & Clarkson, W. 2005, MNRAS, 359, 567Boué, G., Oshagh, M., Montalto, M., & Santos, N. C. 2012, MNRAS,

422, L57Carter, J. A., Agol, E., Chaplin, W. J., et al. 2012, Sci, 337, 556Deck, K. M., & Agol, E. 2015, ApJ, 802, 116Deck, K. M., Agol, E., Holman, M. J., & Nesvorný, D. 2014, ApJ, 787, 132Fabrycky, D. C., Lissauer, J. J., Ragozzine, D., et al. 2014, ApJ, 790, 146Goodman, J., & Weare, J. 2010, CAMCoS, 5, 65Hadden, S., & Lithwick, Y. 2015, ApJ, submitted (arXiv:1510.02476)Holman, M. J., & Murray, N. W. 2005, Sci, 307, 1288Huber, D., Carter, J. A., Barbier, M., et al. 2013, Sci, 342, 331Jontof-Hutter, D., Ford, E. B., Rowe, J. F., et al. 2016, ApJ, 820, 39Jontof-Hutter, D., Rowe, J. F., Lissauer, J. J., Fabrycky, D. C., & Ford, E. B.

2015, Natur, 522, 321Lithwick, Y., Xie, J., & Wu, Y. 2012, ApJ, 761, 122Miralda-Escudé, J. 2002, ApJ, 564, 1019Morbidelli, A. 2002, Modern Celestial Mechanics : Aspects of Solar System

Dynamics (Boca Raton, FL: CRC Press)Murray, C. D., & Dermott, S. F. 1999, Solar System Dynamics (1st ed;

Cambridge: Cambridge Univ. Press)Nesvorný, D. 2009, ApJ, 701, 1116Nesvorný, D., & Beaugé, C. 2010, ApJL, 709, L44Nesvorný, D., Kipping, D., Terrell, D., et al. 2013, ApJ, 777, 3Nesvorný, D., Kipping, D. M., Buchhave, L. A., et al. 2012, Sci, 336, 1133Nesvorný, D., & Morbidelli, A. 2008, ApJ, 688, 636Nesvorný, D., & Vokrouhlický, D. 2014, ApJ, 790, 58Petigura, E. A., Schlieder, J. E., Crossfield, I. J. M., et al. 2015, ApJ, 811, 102

13

The Astrophysical Journal, 821:96 (13pp), 2016 April 20 Deck & Agol