scispace - formally typeset
Search or ask a question
Journal ArticleDOI

A systematic retrieval analysis of secondary eclipse spectra. i. a comparison of atmospheric retrieval techniques

01 Oct 2013-The Astrophysical Journal (American Astronomical Society)-Vol. 775, Iss: 2, pp 137
TL;DR: In this paper, the authors compare three spectral retrieval methods: optimal estimation, differential evolution Markov chain Monte Carlo, and bootstrap Monte Carlo on a synthetic water-dominated hot Jupiter and find that the three approaches agree for high spectral resolution, high signal-to-noise data expected to come from potential future spaceborne missions, but disagree for low-resolution, low signal tonoise spectra representative of current observations.
Abstract: Exoplanet atmosphere spectroscopy enables us to improve our understanding of exoplanets just as remote sensing in our own solar system has increased our understanding of the solar system bodies. The challenge is to quantitatively determine the range of temperatures and molecular abundances allowed by the data, which is often difficult given the low information content of most exoplanet spectra that commonly leads to degeneracies in the interpretation. A variety of spectral retrieval approaches have been applied to exoplanet spectra, but no previous investigations have sought to compare these approaches. We compare three different retrieval methods: optimal estimation, differential evolution Markov chain Monte Carlo, and bootstrap Monte Carlo on a synthetic water-dominated hot Jupiter. We discuss expectations of uncertainties in abundances and temperatures given current and potential future observations. In general, we find that the three approaches agree for high spectral resolution, high signal-to-noise data expected to come from potential future spaceborne missions, but disagree for low-resolution, low signal-to-noise spectra representative of current observations. We also compare the results from a parameterized temperature profile versus a full classical Level-by-Level approach and discriminate in which situations each of these approaches is applicable. Furthermore, we discuss the implications of our models for the inferred C-to-O ratios of exoplanetary atmospheres. Specifically, we show that in the observational limit of a few photometric points, the retrieved C/O is biased toward values near solar and near one simply due to the assumption of uninformative priors.

Summary (4 min read)

1. INTRODUCTION

  • Thermal emission spectra (∼1–30 μm) of extrasolar planets can tell us about their atmospheric temperatures and compositions (see, e.g., Charbonneau et al.
  • The authors call their three-pronged retrieval approach CHIMERA (CaltecH Inverse ModEling and Retrieval Algorithms).

2.1. The Retrieval Techniques

  • The authors use three different retrieval techniques to infer the compositions and temperatures from a spectrum.
  • The techniques are inherently Bayesian as they attempt to solve the inverse problem by summarizing the full shape of the posterior in terms of the location in parameter space of the maximum likelihood and the uncertainties about that location.
  • The first, and the fastest (least number of forward model calls), of these approaches is OE, the second is the model-dependent BMC, and the third is DEMC.

2.1.1. Optimal Estimation (OE)

  • The OE retrieval approach is well established in the fields of Earth atmosphere remote sensing (Rodgers 1976, 2000; Twomey 1996; Kuai et al. 2013), solar system atmosphere remote sensing (Conrath et al.
  • The basic approach is to minimize a cost function to obtain the maximum a posteriori (MAP) solution.
  • The second term represents the prior uncertainties before making the measurements, which has less influence for higher quality data.
  • This assumption is only valid when the region in phase space over which the forward model can be linearized is broader than the parameter uncertainties.

2.1.2. Model-dependent Bootstrap Monte Carlo (BMC)

  • A common way to more robustly characterize errors is through a Monte Carlo resampling of the data (see, e.g., Press et al.
  • These synthetic data are then refit using, say, OE, and the resulting best-fit parameter distributions represent the uncertainties.
  • This new realization is then fit and the process is repeated many times.
  • The approach the authors take is similar, but rather than generate a new spectrum using the residual, they simply take the best fit, from OE, and then resample each point by drawing it from a normal distribution with a mean given by the best-fit value and the width given by the data error bar for that point.

2.1.3. Differential Evolution Markov Chain Monte Carlo (DEMC)

  • The MCMC approach has revolutionized parameter estimation and error analysis in many fields.
  • If the proposed state (xp) has an improved likelihood over the current state, then that state is kept (xi+1 = xp) and a new proposal is made from that location.
  • Rather than standard MCMC approaches, the authors use an adaptive algorithm known as DEMC (ter Braak 2006; ter Braak & Vrugt 2008).
  • Set the last link in one of the chains to the best-fit state vector obtained in step 1.
  • Repeat this process for the other Nchains − 1 chains.

2.2. The Forward Model

  • The forward model, F(x), is the most important part of any retrieval algorithm.
  • In the case of atmospheric retrieval, the forward model takes temperatures and compositions and generates a model spectrum.
  • On that note, the authors use the HITEMP database (Rothman et al. 2010) to compute the tabulated cross sections for CO2, CO, and H2O and the STDS database for CH4 (Wenger & Champion 1998).
  • The authors feel this is appropriate for two reasons.

2.2.1. Parameterized versus Level-by-Level Temperature Profile

  • The authors employ two approaches to retrieve the temperature profiles.
  • The Level-by-Level approach is only appropriate when the information content of the spectra is sufficiently high such that the addition of the Nlev additional parameters is justified.
  • The second temperature profile retrieval approach makes use of a parameterization.
  • This τ–P mapping assumes a linear relation between the optical depth and pressure, or a pressure-independent IR opacity.
  • The authors can think of the prior as an artificial set of “data” on which the retrieval (all retrieval approaches) can rely when the measurements are insufficient to constrain a given parameter.

3.1. Synthetic Observations

  • The authors create a generic hydrogen-dominated hot Jupiter planet and derive its emission spectrum in three different observing scenarios.
  • Figure 2 shows the model atmosphere and spectrum of the synthetic planet.
  • The thermal contribution functions indicate that the emission from shorter wavelengths comes from deeper layers in the atmosphere, and regions of high opacity tend to push the emission to higher altitudes.
  • The size of the error bars are representative of typical errors of IRAC observations (e.g., Machalek et al. 2009).
  • The second observational scenario is a multi-instrument case combining both Spitzer photometry, ground-based photometry, and Hubble WFC3 spectra .

3.2. The Prior

  • As mentioned in Section 2.1.1, the prior is important when the spectral information content is limited.
  • The authors choose extremely broad Gaussian priors in order to mitigate the influence they have on the retrievals.
  • This value could go higher in the presence of very low emissivity and redistribution.
  • Generally the opacity may have a pressure dependence, and hence solving self-consistently for this constant-with-altitude opacity would not actually recover the pressure dependence of this opacity.
  • The prior temperature profile distributions are reconstructed by propagating the Gaussian prior probability distributions (including the above limits) of κIR, γv1 , γv2 , β, and α in Table 2 through Equations (13)–(16).

3.3. Results from the Parameterized Temperature Profile

  • The authors apply the three retrieval techniques to the three synthetic observations in Figure 3 under the radiative equilibrium temperature profile parameterization.
  • The temperature prior in Figure 4 is used in all three techniques for all three retrieval cases.
  • The authors must be careful in interpreting the confidence interval values when the posteriors extend to the imposed upper and lower limits, especially when those limits are somewhat arbitrary.
  • The second and third rows show the fits from the BMC and DEMC, respectively.
  • In the following subsections, the authors summarize posteriors for the gas compositions, temperatures, and C to O ratios for each observational scenario.

3.3.1. Gas Abundance Retrievals

  • The gas mixing ratio retrieval results are summarized in Figures 6 and 7.
  • The first column of Figure 6 shows the marginalized gas posteriors for the broadband observational scenario and the top set of panels of Figure 7 show the correlations amongst the gases.
  • CH4 has a better defined upper edge than CO2 because both the 3.6 μm and 8 μm channels overlap with the strongest methane absorption bands.
  • BMC underestimates the uncertainties in all species with the exception of water.
  • The differences in the 1σ uncertainties derived from OE are less than ∼10% than the uncertainties derived from DEMC.

3.3.2. Temperature Profile Retrievals

  • The marginalized posteriors for the five parameters that govern the shape of the temperature profile for each observational scenario are shown in Figure 8.
  • The dark and light red swaths show the 1σ and 2σ bounds on the reconstructed temperature profiles.
  • Again, the BMC approach completely underestimates the error when compared with the other two approaches because of its inability to fully characterize the posterior outside of a small region of phase space localized around the OE original best fit.
  • OE and BMC underestimate the temperature uncertainties relative to DEMC at 100 mbar, but the OE and DEMC have reasonable agreement over the entire profile.
  • This correlation is prevalent in all three observational scenarios, as even the broadband points are strongly effected by water vapor absorption .

3.3.3. C to O Ratios

  • Determination of the C-to-O ratios of explanatory atmospheres is critical to the understanding of their atmospheric chemistry (Lodders & Fegley 2002; Moses et al. 2011) and formation environments (Öberg et al. 2011; Madhusudhan et al. 2011).
  • Before inspecting the posteriors derived from CHIMERA, the authors find it illustrative to investigate the prior.
  • Upon propagating the Gaussian priors (with the limits) of the gases through Equation (19), the authors obtain the C/O prior in Figure 10.
  • The authors would expect then that one gas will have a larger abundance than the other three 25% of the time.
  • Since the OE best-fit gas abundances are very near truth , the BMC posteriors, which are highly localized about the OE best-fit parameters, will overemphasize the C/O derived by that best fit.

3.4. Results from the Level-by-Level Temperature Profile

  • The Level-by-Level temperature profile approach attempts to determine the temperature for each model layer.
  • The uncertainties in temperature at 100 mbar are reduced to ±177 K, though smaller uncertainties are achieved at deeper levels due to the addition of the WFC3 data which probe deeper atmospheric levels.
  • This is why the averaging kernel profile peaks at a deeper level.
  • As before, in both cases the BMC approach underestimates the temperature uncertainties relative to the OE derived uncertainties.
  • For this, the authors investigate the effect of three different temperature priors (different prior profiles, xa, but same widths, Sa) on the retrieved profiles and check to see if they are consistent with the estimated errors .

4. DISCUSSION AND CONCLUSIONS

  • The authors have developed a new statistically robust suite of exoplanet atmospheric retrieval algorithms known as CHIMERA.
  • In the high signal-to-noise and high spectral resolution regime, both the BMC and OE methods provide reasonable parameter uncertainties.
  • The temperature profiles and corresponding uncertainties can only be trusted for the region over which the thermal emission contribution functions peak, typically between a few bars and a few mbar .
  • The authors also thank John Johnson and Jonathan Fortney for useful conversations.

Did you find this useful? Give us your feedback

Figures (17)

Content maybe subject to copyright    Report

The Astrophysical Journal, 775:137 (22pp), 2013 October 1 doi:10.1088/0004-637X/775/2/137
C
2013. The American Astronomical Society. All rights reserved. Printed in the U.S.A.
A SYSTEMATIC RETRIEVAL ANALYSIS OF SECONDARY ECLIPSE SPECTRA. I.
A COMPARISON OF ATMOSPHERIC RETRIEVAL TECHNIQUES
Michael R. Line
1
, Aaron S. Wolf
1
, Xi Zhang
1
, Heather Knutson
1
, Joshua A. Kammer
1
, Elias Ellison
2
,
Pieter Deroo
3
, Dave Crisp
3
, and Yuk L. Yung
1
1
Division of Geological and Planetary Sciences, California Institute of Technology, Pasadena, CA 91125, USA; mrl@gps.caltech.edu
2
Flintridge Preparatory School, La Ca
˜
nada, CA 91011, USA
3
Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA 91109-8099, USA
Received 2013 April 19; accepted 2013 August 19; published 2013 September 16
ABSTRACT
Exoplanet atmosphere spectroscopy enables us to improve our understanding of exoplanets just as remote sensing in
our own solar system has increased our understanding of the solar system bodies. The challenge is to quantitatively
determine the range of temperatures and molecular abundances allowed by the data, which is often difficult given
the low information content of most exoplanet spectra that commonly leads to degeneracies in the interpretation. A
variety of spectral retrieval approaches have been applied to exoplanet spectra, but no previous investigations have
sought to compare these approaches. We compare three different retrieval methods: optimal estimation, differential
evolution Markov chain Monte Carlo, and bootstrap Monte Carlo on a synthetic water-dominated hot Jupiter. We
discuss expectations of uncertainties in abundances and temperatures given current and potential future observations.
In general, we find that the three approaches agree for high spectral resolution, high signal-to-noise data expected
to come from potential future spaceborne missions, but disagree for low-resolution, low signal-to-noise spectra
representative of current observations. We also compare the results from a parameterized temperature profile versus
a full classical Level-by-Level approach and discriminate in which situations each of these approaches is applicable.
Furthermore, we discuss the implications of our models for the inferred C-to-O ratios of exoplanetary atmospheres.
Specifically, we show that in the observational limit of a few photometric points, the retrieved C/Oisbiasedtoward
values near solar and near one simply due to the assumption of uninformative priors.
Key words: methods: data analysis methods: statistical planets and satellites: atmospheres radiative transfer
Online-only material: color figures
1. INTRODUCTION
Thermal emission spectra (1–30 μm) of extrasolar plan-
ets can tell us about their atmospheric temperatures and com-
positions (see, e.g., Charbonneau et al. 2005; Tinetti et al.
2007, 2010a; Grillmair et al. 2007, 2008; Swain et al. 2009a,
2009b; Madhusudhan & Seager 2009; Stevenson et al. 2010;
Madhusudhan et al. 2011; Lee et al. 2012; Line et al. 2012). At
the moment, these observations come in two types, broadband
photometry mainly from the Spitzer Space Telescope (see, e.g.,
Knutson et al. 2010) and ground-based instruments (Croll et al.
2010; Anderson et al. 2010; Gibson et al. 2010; Deming et al.
2012; Gillon et al. 2012), as well as higher resolution spectra
such as Hubble Space Telescope (HST) Wide Field Camera 3
(WFC3; Berta et al. 2012; Swain et al. 2012; Deming et al.
2013) and Near-Infrared Camera and Multi-Object Spectrom-
eter (Swain et al. 2009a, 2009b; Tinetti et al. 2010a;Gibson
et al. 2011; Crouzet et al. 2012). From these observations,
signatures of a variety of molecules have been detected in-
cluding H
2
O, CH
4
, CO, and CO
2
(Tinetti et al. 2007;Swain
et al. 2009a, 2009b; Tinetti et al. 2010a), although the ro-
bustness of some of these detections have recently been called
into question (Gibson et al. 2011). These same data have been
used to infer the presence of atmospheric temperature inver-
sions for a subset of hot Jupiters (e.g., Burrows et al. 2007;
Knutson et al. 2008, 2010; Fortney et al. 2008; Madhusudhan &
Seager 2009).
While the above studies have given us insight into the nature
of these planetary atmospheres, very few have focused on the
uncertainties in temperatures and compositions. Until relatively
recently (Madhusudhan & Seager 2009; Madhusudhan et al.
2011; Lee et al. 2012; Line et al. 2012), most compositions and
temperatures, and thus the subsequent conclusions, were de-
termined through self-consistent forward modeling approaches
that only explore a few potential solutions without a well-defined
characterization of the uncertainty distributions of the physical
parameters (e.g., Burrows et al. 2005, 2007; Fortney et al. 2005).
Furthermore, some self-consistent solutions make physical as-
sumptions that may not necessarily be valid in exoplanetary
atmospheres such as the assumption of thermochemical equi-
librium gas concentrations or radiative–convective temperature
structures (that is, they may ignore other potentially important
processes such as vertical mixing, photochemistry, zonal winds,
etc.). Additionally, this forward modeling approach is often un-
guided by the data and primarily driven by preconceived notions
of how the atmosphere “should” look (as pointed out by Lee et al.
2012 and Benneke & Seager 2012) with the best solutions being
the few that provide the lowest values of chi-squared.
In order to more rigorously characterize the ranges of
allowable temperatures and compositions, Madhusudhan &
Seager (2009) developed a multidimensional grid search ap-
proach which can fully characterize the uncertainty distributions
for each parameter. Subsequent studies (Madhusudhan et al.
2011; Benneke & Seager 2012) used the more sophisticated
Markov chain Monte Carlo (MCMC) approach to accomplish
this goal. However, such approaches require the computation
of many millions of models in order to fully characterize the
parameter uncertainties which may not be feasible for more so-
phisticated forward models with many free parameters. In order
to remedy this problem, Lee et al. (2012) and Line et al. (2012)
1

The Astrophysical Journal, 775:137 (22pp), 2013 October 1 Line et al.
used the much faster optimal estimation (OE; e.g., Rodgers
2000) approach to estimate the error distributions of each pa-
rameter. This approach is much faster due to the assumption
that the parameter error distributions are Gaussian. However,
this Gaussian assumption may result in an incorrect estimate of
the error distributions (Benneke & Seager 2012).
The goals of this paper are to first understand the composition
and temperature uncertainty distributions for different degrees
of observational quality, and second to understand how those
derived uncertainty distributions differ between the two fun-
damental parameter estimation approaches, OE and MCMC.
This investigation represents the first direct comparison and
synthesis of these retrieval approaches as applied to exoplanet
atmospheres. A secondary goal is to understand how the de-
rived composition uncertainties propagate into the C/O uncer-
tainty. We accomplish these goals by comparing three different
retrieval algorithms: OE, a new MCMC algorithm known as
differential evolution Markov chain Monte Carlo (DEMC), and
the model-dependent bootstrap Monte Carlo (BMC) approach.
This investigation is analogous to the investigation carried out
by Ford (2005) on radial velocity data. First, we will describe
the three different retrieval techniques as well as our forward
model in Section 2. We call our three-pronged retrieval ap-
proach CHIMERA (CaltecH Inverse ModEling and Retrieval
Algorithms). Second, we compare the three spectral retrieval
methods on different synthetic spectral data sets of varying
observational quality in order to assess the robustness of the
error estimations from each approach in Section 3. We will also
compare the parameterized temperature profile approach (e.g.,
Madhusudhan & Seager 2009; Line et al. 2012) with the Level-
by-Level profile approach (Lee et al. 2012). Finally, we will
discuss the implications of these uncertainties for the estimated
C to O ratios.
2. METHODS
In this section, we describe the retrieval techniques, the
forward model, and the parameterizations we use to retrieve the
temperatures and compositions from thermal emission spectra.
2.1. The Retrieval Techniques
We use three different retrieval techniques to infer the
compositions and temperatures from a spectrum. The techniques
are inherently Bayesian as they attempt to solve the inverse
problem by summarizing the full shape of the posterior in terms
of the location in parameter space of the maximum likelihood
and the uncertainties about that location. The first, and the fastest
(least number of forward model calls), of these approaches is
OE, the second is the model-dependent BMC, and the third is
DEMC.
2.1.1. Optimal Estimation (OE)
The OE retrieval approach is well established in the fields
of Earth atmosphere remote sensing (Rodgers 1976, 2000;
Twomey 1996; Kuai et al. 2013), solar system atmosphere
remote sensing (Conrath et al. 1998; Irwin et al. 2008; Nixon
et al. 2007; Fletcher et al. 2007; Greathouse et al. 2011), and
recently exoplanet atmosphere remote sensing (Lee et al. 2012;
Line et al. 2012). The basic approach is to minimize a cost
function to obtain the maximum a posteriori (MAP) solution.
Using Bayes theorem and the assumption that the data likelihood
and the prior are Gaussian, one can derive the following cost
function (or log likelihood):
J (x) = (y F(x))
T
S
1
e
(y F(x)) + (x x
a
)
T
S
1
a
(x x
a
), (1)
where y is the set of n observations, x is the set of m parameters
which we wish to retrieve or the state vector, F(x)istheforward
model that maps the state vector onto the observations (described
in Section 2.2), and S
e
is the n ×n data error covariance matrix
(typically off diagonal terms are zero and the diagonal elements
are the square of the 1σ errors of the observations). x
a
is the a
priori state vector and S
a
is the m×m a priori covariance matrix.
The first term in Equation (1) is simply “chi-squared” and the
second term represents the prior knowledge of the parameter
distribution before we make the observations. For high-quality
observations the second term is generally not important as
most of the information in constraining the state vector comes
from the observations. For low-quality observations it is just
the opposite. Following Irwin et al. (2008), we minimize
Equation (1) with Newton’s iteration method given by
x
i+1
= x
a
+ S
1
a
K
T
i
K
i
S
1
a
K
T
i
+ S
1
e
(F(x) y K
i
(x
a
x
i
)), (2)
where i is the iteration index and K
i
is the Jacobian matrix at i
(K
nm
= ∂F
n
/∂x
m
). Rather than taking the full Newton step, we
damp the solution with
x
i+1
= x
i
+
x
i+1
x
i
1 + ζ
, (3)
where ζ is the damping parameter. At each iteration, we evaluate
J(x
i+1
) and J(x
i+1
). If the latter is smaller, we set the state vector
for the next iteration to x
i+1
and decrease ζ by 0.3. Otherwise, we
keep increasing ζ by a factor of 10 and re-evaluate Equations (1)
and (3) until J(x
i+1
) becomes less than J(x
i+1
). Convergence is
achieved when J changes by less than 1×10
6
from the previous
iteration, which typically occurs after 10 s of iterations. The
resulting state vector is the MAP solution, or the “best fit.
Assuming that the posterior is normal, which is achieved by
linearizing the forward model about the best-fit solution, the
uncertainties on the state vector parameters are given by the
posterior covariance matrix:
ˆ
S =
K
T
S
1
e
K + S
1
a
1
. (4)
Again, this matrix represents a multi-dimensional normal dis-
tribution (see Rodgers 2000 for the derivation). The diagonal
elements are the square of the marginalized errors, whereas
the off diagonal terms describe the correlations/degeneracies
amongst the parameters. The first term, K
T
S
1
e
K, represents the
uncertainties due to the measurement errors. This term uses the
local gradient information to estimate the parameter uncertain-
ties. The second term represents the prior uncertainties before
making the measurements, which has less influence for higher
quality data. Again, the major assumption in Equations (1)
and (4) is that the posterior for each parameter is Gaussian.
This assumption is only valid when the region in phase space
over which the forward model can be linearized is broader than
the parameter uncertainties. However, it is this assumption that
allows this approach to be extremely fast requiring only tens
of forward model calls, which given the speed of our forward
model (5 s), results in a full retrieval in only a few minutes. As
we shall see in Section 3, this assumption is valid for data that
is of high resolution and signal to noise, but breaks down for
2

The Astrophysical Journal, 775:137 (22pp), 2013 October 1 Line et al.
low-resolution, low signal-to-noise data. In order to ensure that
the global minimum of Equation (1) is found, multiple start-
ing guesses are used. They generally all converge to the same
solution.
2.1.2. Model-dependent Bootstrap Monte Carlo (BMC)
A common way to more robustly characterize errors is
through a Monte Carlo resampling of the data (see, e.g., Press
et al. 1995, Chapter 15.6; Ford 2005, Section 4.2) in which
many thousands of realizations of the original data (in our
case, the spectra) are created using the uncertainties from the
original data set. These synthetic data are then refit using, say,
OE, and the resulting best-fit parameter distributions represent
the uncertainties. There are multiple ways of generating the
synthetic data realizations. The most common way is the residual
resampling approach in which data realizations are created
by adding the randomly shuffled residual between the best-fit
model and the data back to the original best-fit model. This new
realization is then fit and the process is repeated many times.
The approach we take is similar, but rather than generate a new
spectrum using the residual, we simply take the best fit, from
OE, and then resample each point by drawing it from a normal
distribution with a mean given by the best-fit value and the width
given by the data error bar for that point. We chose this approach
over the residual resampling approach because sparse coverage
spectra, like those from broadband observations, have virtually
no residual as they can be fit perfectly due to the greater number
of parameters than data points. We typically generate 1000
spectra realizations that are then refit by OE to obtain the state
vector parameter distributions.
2.1.3. Differential Evolution Markov Chain Monte Carlo (DEMC)
The MCMC approach has revolutionized parameter estima-
tion and error analysis in many fields. It is routinely used
in radial velocity (Ford 2005) and transit light curve (e.g.,
Eastman et al. 2013) error analysis. Results from a well con-
verged MCMC analysis can generally be considered as the best
possible representation of the parameter uncertainties. Recently,
this approach has been applied to the exoplanet atmosphere re-
trieval problem (Madhusudhan et al. 2011; Benneke & Seager
2012). Unlike OE, MCMC approaches make no assumptions
about the shape of the posterior, but rather evaluate the posterior
with millions of samples.
The basic approach of MCMC is to sample the posterior
through a random walk process. The random walk is carried
out by drawing states from some proposal distribution and
evaluating whether or not the proposed state has an increased
likelihood over the previous. Typically, the proposal distribution
is a normal distribution with a mean given by the current state
in the chain (x
i
) and a user-defined width to achieve a particular
acceptance rate (Gibbs sampling or Metropolis–Hastings). If the
proposed state (x
p
) has an improved likelihood over the current
state, then that state is kept (x
i+1
= x
p
) and a new proposal is
made from that location. If the proposal state has not improved
the likelihood, then that state is either rejected or accepted with
some probability. This previous state-dependent random walk
constitutes a Markov chain. Given enough samples, this Markov
chain will converge to the target posterior (see Ford 2005 for a
more detailed explanation).
Rather than standard MCMC approaches, we use an adaptive
algorithm known as DEMC (ter Braak 2006; ter Braak &
Vrugt 2008). The purpose of this approach is to obtain more
appropriate proposal states by identifying the proper scale and
orientation of the current estimate of the posterior. This scale
and orientation information comes from the chain history. This
approach gives a more efficient probing method for highly
correlated parameter spaces and yields improved convergence
rates. Our DEMC procedure is as follows.
1. Apply the OE technique to the measurements to obtain the
best-fit state vector and posterior covariance matrix,
ˆ
S.This
step provides an initial estimate of the posterior.
2. Initialize N
init
links (x
i=0N
init
) in each of the N
chains
(typ-
ically three chains, more chains will slow convergence)
independent chains (arrays) by randomly drawing state vec-
tors from the multivariate normal described by the posterior
covariance matrix from step 1. Set the last link in one of
the chains to the best-fit state vector obtained in step 1.
This step provides a good starting history from which our
initial proposal states can be drawn. Combine each of the
independent chains into one long chain that composes the
history, X
history
.
3. Evaluate the cost function, J, in Equation (1) for the last link
in each of the chains. If using a flat prior, ignore the second
term. Again, this is simply the equivalent of evaluating chi-
squared.
4. Draw two random numbers, R
1
and R
2
, between zero and
N
chains
×i, where i is the current state in the chain. Initially,
i = N
init
. Evaluate the proposed jump state given by
x
p
= x
i
+ γ (x
R1
x
R2
)+e, (5)
where x
R
1
and x
R
2
are the states from different points in
the chain history, X
history
. γ is a scale factor typically set to
2.38/
(2 m) (ter Braak 2006), where m is the number of
parameters. This factor is meant to give acceptance rates of
0.23 for large m. e is a vector drawn from a multivariate
normal distribution with a small variance relative to the
chain variance in order to introduce a small amount of
additional randomness. Repeat this process for the other
N
chains
1 chains.
5. Evaluate the Metropolis (Metropolis et al. 1953) ratio,
r = P (x
p
)/P (x
i
) = e
1
2
(J(x
p
)J(x
i
))
.Ifr is larger than 1,
set x
i+1
= x
p
and if it is smaller only accept if it is larger
than a random number between 1 and 0. Otherwise, do
not update the chain, set x
i+1
= x
i
. Repeat for the other
N
chains
1 chains. Add the updated links in all N
chains
to
X
history
.
6. Repeat steps 4 and 5 until convergence is met. Convergence
can be determined by looking at the trace plots of X
history
for
each parameter or by using the Gelman–Rubin statistic on
the set of N
chains
chains. For this we use the algorithm from
Eastman et al. (2013) which requires the Gelman–Rubin
statistic to be less than 1.01 and the number of independent
draws to be greater than 1000 for each parameter. Conver-
gence typically occurs in less than 10
5
links in each of the
N
chains
for a total of N
chains
× 10
5
links, which, given the
speed of our forward model, takes 5 days for a typical
run. This is about an order of magnitude less than parallel
tempering or pure Metropolis–Hastings.
2.2. The Forward Model
The forward model, F(x), is the most important part of any re-
trieval algorithm. It is what maps the state vector of retrievable
parameters onto the observations. In the case of atmospheric
retrieval, the forward model takes temperatures and composi-
tions and generates a model spectrum. Our particular forward
3

The Astrophysical Journal, 775:137 (22pp), 2013 October 1 Line et al.
model numerically solves the planet-parallel thermal infrared
(IR) radiation problem for an absorbing, emitting atmosphere
(we neglect scattering). We first divide the atmosphere into N
lev
discretized atmospheric layers. The absorption optical depth for
the kth gas in the zth layer at wavelength λ is
Δτ
k,z,λ
= f
k,z
σ
k,z,λ
ΔP
z
μ
atm
g
, (6)
where f
k,z
is the volume mixing ratio of the kth gas in the zth
layer, σ
k,z,λ
is the absorption cross section per molecule of the kth
gas in zth layer at wavelength λ, ΔP
z
is pressure thickness of the
zth slab, μ
atm
is the mean molecular weight of the atmosphere,
and g is the gravity. The absorption cross sections are pre-
computed on a 1 cm
1
wavenumber grid at 20 evenly spaced
temperature and log-pressure points from 500–3000 K and
50–10
6
bars, respectively (similar to Sharp & Burrows 2007).
The cross sections for each wavelength on the pre-computed grid
are interpolated to the atmospheric temperatures and pressures
in the zth slab. To compute the total slab optical depth, we sum
the contribution from each gas to obtain
Δτ
z,λ
=
N
gas
k=1
Δτ
k,z,λ
. (7)
Upon computing the optical depths at each level, we can now
solve for the upwelling irradiance with
I
λ
=
N
lev
z=0
B
λ
(T
z
)e
N
lev
j=z
Δτ
j,λ
Δτ
z,λ
, (8)
where N
lev
is the number of atmospheric levels and B
λ
(T
z
)is
the Planck function at wavelength λ and temperature in the zth
slab. We use 90 atmospheric layers to compute the upwelling
flux.
An important part of the forward model when using the OE
approach is the computation of the Jacobian, or the sensitivity to
the upwelling irradiance with respect to the desired retrievable
parameters. When possible, it is preferable that the Jacobian
be calculated analytically for both improvements in speed
and in accuracy. We are interested in the retrieval of both
abundances and temperatures so we must compute the Jacobian
with respect to both the abundances and temperatures. We
make the assumption of vertically uniform gas mixing ratios
throughout the atmosphere, and hence f
k,z
is independent of z.
We now differentiate Equation (8) with respect to the uniform
gas mixing ratios for each gas f
k
to obtain
∂I
λ
∂f
k
=
N
lev
z=0
B
λ
(T
z
)e
N
lev
j=z
Δτ
j,λ
Δτ
k,z,λ
f
k
N
lev
z=0
B
λ
(T
z
)e
N
lev
j=z
Δτ
j,λ
Δτ
z,λ
N
lev
j=z
Δτ
k,j,λ
f
k
. (9)
The first term is due to the changing emissivity of the emitting
slab and the second term is how the change in transmittance
affects the upwelling irradiance.
The sensitivity of the irradiance to a change in temperature
in the zth slab is given by
∂I
λ
∂T
z
=
e
N
lev
j=z+1
Δτ
j,λ
e
N
lev
j=z
Δτ
j,λ
∂B
λ
(T
z
)
∂T
z
. (10)
This equation is similar to Equation (14) in Irwin et al. (2008)
but we have neglected the first and last terms in their formula as
they are small.
Since the observations are reported as the ratio of the planet
flux to the stellar flux and not the irradiance, we perform a
disk integration of Equations (8)–(10) using four point Gaussian
quadrature and then divide by an interpolated PHOENIX stellar
flux grid model (Allard et al. 2000).
We include only CH
4
,CO
2
,CO,H
2
O, H
2
, and He in our
model. H
2
and He are fixed in our models at thermochemically
justifiable abundances. The exact abundances of these species
are not critical as the sensitivity of the spectrum to H
2
and
He is minimal. We retrieve only CH
4
,CO
2
, CO, and H
2
O. We
choose these species because they are the most spectroscopically
active and abundant species. Admittedly, we could/should
include every possible atmospheric constituent but this would
be unwieldy and reliable high-temperature absorption line lists
only exist for a few. On that note, we use the HITEMP
database (Rothman et al. 2010) to compute the tabulated cross
sections for CO
2
, CO, and H
2
O and the STDS database for
CH
4
(Wenger & Champion 1998). Below 1.7 μmforCH
4
,
we simply use the HITRAN (Rothman et al. 2009) database
for lack of anything better (to the best of our knowledge).
We use the Borysow et al. (2001), Borysow (2002), and
Jørgensen et al. (2000) databases for the computation of the
H
2
–H
2
/He collision induced opacities. The Reference Forward
Model (http://www.atm.ox.ac.uk/RFM/) was used to compute
the tabulated cross sections from the line strength databases. We
have validated our forward model through a detailed comparison
with the Oxford NEMESIS group (e.g., Lee et al. 2012) and our
results agree to better than 5% (see Figure 1).
An additional component of the forward model is the instru-
mental function used to convolve with the high-resolution model
spectrum. For the broadband points we simply integrate the flux
from the high-resolution model spectrum with the appropriate
filter function for that point. When fitting higher resolution ob-
servations, the instrumental function is assumed to be a Gaussian
(valid for grating spectrometers) in wavelength with an FWHM
determined by observations.
Now that we have a well-defined forward model we can define
our state vector. Again, we wish to retrieve the abundances of
CH
4
,CO
2
, CO, and H
2
O and the temperature profile. More
specifically, we choose to retrieve the log of the abundances as
they can vary by orders of magnitude and to prevent negative
mixing ratios. Our state vector is given by
x = [log(f
H
2
O
), log(f
CH
4
), log(f
CO
), log(f
CO
2
),T]
T
, (11)
where the f
k
s are all assumed constant with altitude. We feel this
is appropriate for two reasons. First, vertical mixing will smooth
out the mixing ratio profiles over the thermal IR photosphere
(Line et al. 2010, 2011; Moses et al. 2011), and second, current
observations simply do not have the information content to
warrant the retrieval of vertical mixing ratio information (see
Lee et al. 2012). In the next section, we describe how to go
about retrieving the temperature profile.
2.2.1. Parameterized versus Level-by-Level Temperature Profile
We employ two approaches to retrieve the temperature
profiles. The first, and the most commonly used in Earth and
solar system atmosphere retrieval problems, is the Level-by-
Level approach. This is the approach used in Lee et al. (2012).
The second is a parameterized temperature profile approach
4

The Astrophysical Journal, 775:137 (22pp), 2013 October 1 Line et al.
NEMESIS Forward Model
CHIMERA Forward Model
CH
4
CO
2
CO
H
2
O
Figure 1. Comparison of the thermal emission spectrum from our forward model (black) with the NEMESIS forward model (red). The temperature–pressure profile
is shown in the inset. For this comparison we assume uniform mixing ratios of 10
4
for CH
4
,CO
2
,CO,andH
2
O. H
2
is set to 0.85 and He is set to 0.15. This planet
is assumed to be hydrogen dominated (mean molecular weight of 2.3 amu) with a radius of 1 R
J
, a gravity of 22 ms
2
, orbiting a 5700 K pure blackbody star with a
radius of 1 R
sun
.
(A color version of this figure is available in the online journal.)
similar to the approach used in Madhusudhan & Seager (2009)
and Line et al. (2012). Each has its advantages and disadvantages
described below.
The Level-by-Level temperature retrieval approach seeks
an estimate of the temperature at each of the N
lev
model
layers. This approach is advantageous in that there are no
pre-conceived assumptions made about how the atmospheric
temperature should be parameterized. If the spectral signal to
noise and resolution are high enough, there is generally enough
sensitivity to obtain information about the temperature in
individual atmospheric layers. However, there is a finite vertical
resolution given the quality of the observations. Typically,
this resolution is set by the width of the thermal emission
weighting functions and how much they overlap. Generally,
when the spectra are noisy the Level-by-Level approach fits
the noise which results in unphysical structure in the retrieved
temperature profile. This is analogous to fitting a high-degree
polynomial to only a few points. There are ways to smooth
unphysical structure, one of them being to assume a correlation
among the atmospheric layers (Rodgers 2000; Irwin et al. 2008)
implemented through the prior covariance matrix, S
a
, with
S
a,ij
= (S
a,ii
S
a,jj
)
1/2
e
−|ln(P
i
/P
j
)|
h
. (12)
Here P
i
and P
j
are the pressures at the ith and jth levels,
respectively, and h is the correlation length that controls the
level of smoothing. The correlation length can be thought of
as the number of scale heights over which the temperatures
are correlated. For our simulations we choose h = 7asthis
provides a sufficient level of detail without producing unphysical
oscillations. When using this approach, our state vector is
exactly as it is in Equation (11) with T being an N
lev
vector
of temperatures at each level. The Level-by-Level approach is
only appropriate when the information content of the spectra
is sufficiently high such that the addition of the N
lev
additional
parameters is justified. For most current exoplanet spectra, this
is an invalid approach.
The second temperature profile retrieval approach makes use
of a parameterization. This approach is advantageous when the
information content of a spectrum is low as the number of free
variables is greatly reduced. However, the parameterization does
force the retrieved atmospheric temperature structure to conform
only to the profile shapes and physical approximations allowed
by that parameterization. For our particular parameterization,
we assume the atmosphere to be in radiative equilibrium based
upon the analytic radiative equilibrium temperature profile of
Guillot (2010, and others such as Hansen 2008; Heng et al. 2012;
Robinson & Catling 2012). This is the same parameterization
5

Citations
More filters
Journal ArticleDOI
TL;DR: In this article, the authors explore how well spectra from the James Webb Space Telescope (JWST) will likely constrain bulk atmospheric properties of transiting exoplanets.
Abstract: We explore how well spectra from the James Webb Space Telescope (JWST) will likely constrain bulk atmospheric properties of transiting exoplanets. We start by modeling the atmospheres of archetypal hot Jupiter, warm Neptune, warm sub-Neptune, and cool super-Earth planets with atmospheres that are clear, cloudy, or of high mean molecular weight (HMMW). Next we simulate the λ = 1–11 μm transmission and emission spectra of these systems for several JWST instrument modes for single-transit or single-eclipse events. We then perform retrievals to determine how well temperatures and molecular mixing ratios (CH4, CO, CO2, H2O, NH3) can be constrained. We find that λ = 1–2.5 μm transmission spectra will often constrain the major molecular constituents of clear solar-composition atmospheres well. Cloudy or HMMW atmospheres will often require full 1–11 μm spectra for good constraints, and emission data may be more useful in cases of sufficiently high Fp and high Fp/F*. Strong temperature inversions in the solar-composition hot-Jupiter atmosphere should be detectable with 1–2.5+ μm emission spectra, and 1–5+ μm emission spectra will constrain the temperature–pressure profiles of warm planets. Transmission spectra over 1–5+ μm will constrain [Fe/H] values to better than 0.5 dex for the clear atmospheres of the hot and warm planets studied. Carbon-to-oxygen ratios can be constrained to better than a factor of 2 in some systems. We expect that these results will provide useful predictions of the scientific value of single-event JWST spectra until its on-orbit performance is known.

473 citations

Journal ArticleDOI
TL;DR: In this paper, the water abundance in the atmosphere of the 2 MJup short-period exoplanet WASP-43b was determined based on thermal emission and transmission spectroscopy measurements obtained with the Hubble Space Telescope.
Abstract: The water abundance in a planetary atmosphere provides a key constraint on the planet’s primordial origins because water ice is expected to play an important role in the core accretion model of planet formation. However, the water content of the Solar System giant planets is not well known because water is sequestered in clouds deep in their atmospheres. By contrast, short-period exoplanets have such high temperatures that their atmospheres have water in the gas phase, making it possible to measure the water abundance for these objects. We present a precise determination of the water abundance in the atmosphere of the 2 MJup short-period exoplanet WASP-43b based on thermal emission and transmission spectroscopy measurements obtained with the Hubble Space Telescope. We nd the water content is consistent with the value expected in a solar composition gas at planetary temperatures (0:4 3:5 solar at 1 condence). The metallicity of WASP-43b’s atmosphere suggested by this result extends the trend observed in the Solar System of lower metal enrichment for higher planet masses. Subject headings: planets and satellites: atmospheres | planets and satellites: composition | planets and satellites: individual: WASP-43b

432 citations


Cites methods from "A systematic retrieval analysis of ..."

  • ...For the transmission spectrum modeling, we retrieved an effective scale height temperature, a reference pressure at which the fiducial radius is defined, and an opaque gray cloud top pressure, in addition to the molecular abundances (Line et al. 2013a)....

    [...]

  • ...We retrieved the planet’s atmospheric properties by fitting the transmission and dayside emission spectra with the CHIMERA Bayesian retrieval suite (Line et al. 2013b; Line & Yung 2013; Line et al. 2014)....

    [...]

Journal ArticleDOI
TL;DR: The ExoMol database as mentioned in this paper provides extensive line lists of molecular transitions which are valid over extended temperature ranges, including lifetimes of individual states, temperature-dependent cooling functions, Lande g-factors, partition functions, cross sections, k-coefficients and transition dipoles with phase relations.

421 citations


Cites methods from "A systematic retrieval analysis of ..."

  • ...ExoMol data have been incorporated into many radiative transfer and retrieval codes including Tau-REx [13–15], Phoenix [16], NEMESIS [17], CHIMERA [18], HELIOS [19] and HELIOS-r2 [20], ATMO [21–23], ARCiS [24], ARTES [25], Hy- DRA [26], GENESIS [27], petitRADTRANS [28], PLATON [29], VSTAR [30, 31], BART [32] and Pyrat Bay [33]....

    [...]

  • ...ExoMol data have been incorporated into many radiative transfer and retrieval codes including Tau-REx [13–15], Phoenix [16], NEMESIS [17], CHIMERA [18], HELIOS [19] and HELIOS-r2 [20], ATMO [21–23], ARCiS [24], ARTES [25], HyDRA [26], GENESIS [27], petitRADTRANS [28], PLATON [29], VSTAR [30, 31], BART [32] and Pyrat Bay [33]....

    [...]

Journal ArticleDOI
TL;DR: In this article, the authors explore how well James Webb Space Telescope (JWST) spectra will likely constrain bulk atmospheric properties of transiting exoplanets, and they find that the JWST spectra can often constrain the major molecular constituents of clear solar composition atmospheres well.
Abstract: We explore how well James Webb Space Telescope (JWST) spectra will likely constrain bulk atmospheric properties of transiting exoplanets. We start by modeling the atmospheres of archetypal hot Jupiter, warm Neptune, warm sub-Neptune, and cool super-Earth planets with clear, cloudy, or high mean molecular weight atmospheres. Next we simulate the $\lambda = 1 - 11$ $\mu$m transmission and emission spectra of these systems for several JWST instrument modes for single transit and eclipse events. We then perform retrievals to determine how well temperatures and molecular mixing ratios (CH$_4$, CO, CO$_2$, H$_2$O, NH$_3$) can be constrained. We find that $\lambda = 1 - 2.5$ $\mu$m transmission spectra will often constrain the major molecular constituents of clear solar composition atmospheres well. Cloudy or high mean molecular weight atmospheres will often require full $1 - 11$ $\mu$m spectra for good constraints, and emission data may be more useful in cases of sufficiently high $F_p$ and high $F_p/F_*$. Strong temperature inversions in the solar composition hot Jupiter atmosphere should be detectable with $1 - 2.5+$ $\mu$m emission spectra, and $1 - 5+$ $\mu$m emission spectra will constrain the temperature-pressure profiles of warm planets. Transmission spectra over $1 - 5+$ $\mu$m will constrain [Fe/H] values to better than 0.5 dex for the clear atmospheres of the hot and warm planets studied. Carbon-to-oxygen ratios can be constrained to better than a factor of 2 in some systems. We expect that these results will provide useful predictions of the scientific value of single event JWST spectra until its on-orbit performance is known.

351 citations


Cites background or methods from "A systematic retrieval analysis of ..."

  • ...The uncertainties in molecular mixing ratios retrieved from JWST emission spectra (even λ ≥ 5 µm) for single observations are too large to obtain a definitive detection of vertical mixing via the method of Line et al. (2013b)....

    [...]

  • ...The emission spectra are computed using the forward model described in Line et al. (2013a) and subsequent upgrades described in Diamond-Lowe et al. (2014) and Stevenson et al. (2014)....

    [...]

  • ...The transmission spectra are computed with the forward model described in Line et al. (2013b), Swain et al. (2014), and Kreidberg et al. (2014b)....

    [...]

  • ...Several measured flat exoplanet transmission spectra have been reasonably well described with such a parameterization (Line et al. 2013b; Kreidberg et al. 2014a; Knutson et al. 2014)....

    [...]

  • ...…of molecules such CO, CO2, and CH4 in addition to H2O in ∼ 10 exoplanet atmospheres (e.g., Madhusudhan & Seager 2009; Madhusudhan et al. 2011; Lee et al. 2012; Line et al. 2012; Barstow et al. 2013a,b; Benneke & Seager 2012, 2013; Line et al. 2013a, 2014a; Benneke 2015; Waldmann et al. 2015)....

    [...]

Journal ArticleDOI
TL;DR: In this article, a near-infrared transmission spectrum for WASP-12b based on six transit observations with the Hubble Space Telescope/Wide Field Camera 3 (HST-WFC) was reported.
Abstract: Detailed characterization of exoplanets has begun to yield measurements of their atmospheric properties that constrain the planets’ origins and evolution. For example, past observations of the dayside emission spectrum of the hot Jupiter WASP-12b indicated that its atmosphere has a high carbon-tooxygen ratio (C/O > 1), suggesting it had a dierent formation pathway than is commonly assumed for giant planets. Here we report a precise near-infrared transmission spectrum for WASP-12b based on six transit observations with the Hubble Space Telescope/Wide Field Camera 3. We bin the data in 13 spectrophotometric light curves from 0.84 - 1.67 m and measure the transit depths to a median precision of 51 ppm. We retrieve the atmospheric properties using the transmission spectrum and nd strong evidence for water absorption (7 condence). This detection marks the rst high-condence, spectroscopic identication of a molecule in the atmosphere of WASP-12b. The retrieved 1 water volume mixing ratio is between 10 5 10 2 , which is consistent with C/O > 1 to within 2 . However, we also introduce a new retrieval parameterization that ts for C/O and metallicity under the assumption of chemical equilibrium. With this approach, we constrain C/O to 0:5 +0:2 0:3 at 1 and rule out a carbon-rich atmosphere composition (C/O> 1) at > 3 condence. Further observations and modeling of the planet’s global thermal structure and dynamics would aid in resolving the tension between our inferred C/O and previous constraints. Our ndings highlight the importance of obtaining high-precision data with multiple observing techniques in order to obtain robust constraints on the chemistry and physics of exoplanet atmospheres. Subject headings: planets and satellites: atmospheres | planets and satellites: composition | planets and satellites: individual: WASP-12b

320 citations

References
More filters
Journal ArticleDOI
TL;DR: In this paper, a Markov chain Monte Carlo (MCMCMC) technique is used to estimate the uncertainties in the orbital solutions that have been fitted to these observations. But the MCMC technique is not suitable for the high-dimensional parameter spaces necessary for the multiple-planet systems.
Abstract: Precise radial velocity measurements have led to the discovery of ~100 extrasolar planetary systems. We investigate the uncertainty in the orbital solutions that have been fitted to these observations. Understanding these uncertainties will become more and more important as the discovery space for extrasolar planets shifts to longer and longer periods. While detections of short-period planets can be rapidly refined, planets with long orbital periods will require observations spanning decades to constrain the orbital parameters precisely. Already in some cases, multiple distinct orbital solutions provide similarly good fits, particularly in multiple-planet systems. We present a method for quantifying the uncertainties in orbital fits and addressing specific questions directly from the observational data rather than relying on best-fit orbital solutions. This Markov chain Monte Carlo (MCMC) technique has the advantage that it is well suited to the high-dimensional parameter spaces necessary for the multiple-planet systems. We apply the MCMC technique to several extrasolar planetary systems, assessing the uncertainties in orbital elements for several systems. Our MCMC simulations demonstrate that for some systems there are strong correlations between orbital parameters and/or significant non-Gaussianities in parameter distributions, even though the measurement errors are nearly Gaussian. Once these effects are considered, the actual uncertainties in orbital elements can be significantly larger or smaller than the published uncertainties. We also present simple applications of our methods, such as predicting the times of possible transits for GJ 876.

461 citations

Journal ArticleDOI
TL;DR: In this paper, a simple analytical approach inspired by Eddington's approximation for stellar atmospheres was used to derive a relation between temperature and optical depth valid for plane-parallel static grey atmospheres which are both transporting an intrinsic heat flux and receiving an outer radiation flux.
Abstract: The evolution of stars and planets is mostly controlled by the properties of their atmosphere. This is particularly true in the case of exoplanets close to their stars, for which one has to account both for an (often intense) irradiation flux, and from an intrinsic flux responsible for the progressive loss of the inner planetary heat. The goals of the present work are to help understanding the coupling between radiative transfer and advection in exoplanetary atmospheres and to provide constraints on the temperatures of the deep atmospheres. This is crucial in assessing whether modifying assumed opacity sources and/or heat transport may explain the inflated sizes of a significant number of giant exoplanets found so far. I use a simple analytical approach inspired by Eddington's approximation for stellar atmospheres to derive a relation between temperature and optical depth valid for plane-parallel static grey atmospheres which are both transporting an intrinsic heat flux and receiving an outer radiation flux. The model is parameterized as a function of mean visible and thermal opacities, respectively. The model is shown to reproduce relatively well temperature profiles obtained from more sophisticated radiative transfer calculations of exoplanetary atmospheres. It naturally explains why a temperature inversion (stratosphere) appears when the opacity in the optical becomes significant compared to that in the infrared. I further show that the mean equivalent flux (proportional to T4) is conserved in the presence of horizontal advection on constant optical depth levels. This implies with these hypotheses that the deep atmospheric temperature used as outer boundary for the evolution models should be calculated from models pertaining to the entire planetary atmosphere, not from ones that are relevant to the day side or to the substellar point. In these conditions, present-day models yield deep temperatures that are ~1000K too cold to explain the present size of planet HD 209458b. A tenfold increase in the infrared to visible opacity ratio would be required to slow the planetary cooling and contraction sufficiently to explain its size. However, the mean equivalent flux is not conserved anymore in the presence of opacity variations, or in the case of non-radiative vertical transport of energy: The presence of clouds on the night side or a downward transport of kinetic energy and its dissipation at deep levels would help making the deep atmosphere hotter and may explain the inflated sizes of giant exoplanets.

459 citations


"A systematic retrieval analysis of ..." refers methods in this paper

  • ...For our particular parameterization, we assume the atmosphere to be in radiative equilibrium based upon the analytic radiative equilibrium temperature profile of Guillot 2010 (and others such as Hansen 2008; Heng et al. 2012; Robinson & Catling 2012)....

    [...]

Journal ArticleDOI
TL;DR: In this paper, a one-dimensional photochemical and thermochemical kinetics and diffusion model was developed to study the effects of disequilibrium chemistry on the atmospheric composition of hot-Jupiter exoplanets.
Abstract: We have developed a one-dimensional photochemical and thermochemical kinetics and diffusion model to study the effects of disequilibrium chemistry on the atmospheric composition of “hot-Jupiter” exoplanets. Here we investigate the coupled chemistry of neutral carbon, hydrogen, oxygen, and nitrogen species on HD 189733b and HD 209458b and we compare the model results with existing transit and eclipse observations. We find that the vertical profiles of molecular constituents are significantly affected by transport-induced quenching and photochemistry, particularly on the cooler HD 189733b; however, the warmer stratospheric temperatures on HD 209458b help maintain thermochemical equilibrium and reduce the effects of disequilibrium chemistry. For both planets, the methane and ammonia mole fractions are found to be enhanced over their equilibrium values at pressures of a few bar to less than an mbar due to transport-induced quenching, but CH4 and NH3 are photochemically removed at higher altitudes. Disequilibrium chemistry also enhances atomic species, unsaturated hydrocarbons (particularly C2H2), some nitriles (particularly HCN), and radicals like OH, CH3, and NH2. In contrast, CO, H2O, N2, and CO2 more closely follow their equilibrium profiles, except at pressures 1 μbar, where CO, H2O, and N2 are photochemically destroyed and CO2 is produced before its eventual high-altitude destruction. The enhanced abundances of CH4, NH3, and HCN are expected to affect the spectral signatures and thermal profiles of HD 189733b and other relatively cool, transiting exoplanets. We examine the sensitivity of our results to the assumed temperature structure and eddy diffusion coefficients and discuss further observational consequences of these models.

455 citations

Journal ArticleDOI
TL;DR: This approach extends the practical applicability of DE-MC with a snooker updater and is shown to be about 5–26 times more efficient than the optimal Normal random walk Metropolis sampler for the 97.5% point of a variable from a 25–50 dimensional Student t3 distribution.
Abstract: Differential Evolution Markov Chain (DE-MC) is an adaptive MCMC algorithm, in which multiple chains are run in parallel. Standard DE-MC requires at least N=2d chains to be run in parallel, where d is the dimensionality of the posterior. This paper extends DE-MC with a snooker updater and shows by simulation and real examples that DE-MC can work for d up to 50---100 with fewer parallel chains (e.g. N=3) by exploiting information from their past by generating jumps from differences of pairs of past states. This approach extends the practical applicability of DE-MC and is shown to be about 5---26 times more efficient than the optimal Normal random walk Metropolis sampler for the 97.5% point of a variable from a 25---50 dimensional Student t 3 distribution. In a nonlinear mixed effects model example the approach outperformed a block-updater geared to the specific features of the model.

451 citations


"A systematic retrieval analysis of ..." refers methods in this paper

  • ...Rather than standard MCMC approaches, we use an adaptive algorithm known as differential evolution Markov chain Monte Carlo (DEMC) (ter Braak 2006; ter Braak & Vrugt 2008)....

    [...]

Journal ArticleDOI
TL;DR: In this article, the transmission spectrum of the super-Earth exoplanet GJ1214b was estimated to be at between 1.1 and 1.7 m in the transiting phase.
Abstract: Capitalizing on the observational advantage oered by its tiny M dwarf host, we present HST/WFC3 grism measurements of the transmission spectrum of the super-Earth exoplanet GJ1214b. These are the rst published WFC3 observations of a transiting exoplanet atmosphere. After correcting for a ramp-like instrumental systematic, we achieve nearly photon-limited precision in these observations, nding the transmission spectrum of GJ1214b to be at between 1.1 and 1.7 m. Inconsistent with a cloud-free solar composition atmosphere at 8:2 , the measured achromatic transit depth most likely implies a large mean molecular weight for GJ1214b’s outer envelope. A dense atmosphere rules out bulk compositions for GJ1214b that explain its large radius by the presence of a very low density gas layer surrounding the planet. High-altitude clouds can alternatively explain the at transmission spectrum, but they would need to be optically thick up to 10 mbar or consist of particles with a range of sizes approaching 1 m in diameter. Subject headings: planetary systems: individual (GJ 1214b) | eclipses | techniques: spectroscopic

430 citations


"A systematic retrieval analysis of ..." refers background in this paper

  • ...Upon inspecting Figure 7 we find that the WFC3 data combined with the ground based photometric points trims the 2σ tail but does little to improve the 1σ uncertainties....

    [...]

  • ...The uncertainties in temperature at 100 mbars are reduced to ±177 K, though smaller uncertainties are achieved at deeper levels due to the addition of the WFC3 data which probe deeper atmospheric levels....

    [...]

  • ...This is primarily due to the limited spectral coverage provided by the red grism on WFC3, which spans the wavelengths from 1.2-1.6 µm....

    [...]

  • ...At the moment, these observations come in two types, broadband photometry mainly from the Spitzer Space Telescope (see e.g., Knutson et al. 2010) and ground-based instruments (Croll et al. 2010; Anderson et al. 2010; Gibson et al. 2010; Deming et al. 2012; Gillon et al. 2012 ), as well as higher resolution spectra such as Hubble Space Telescopes Wide Field Camera 3 (WFC3) (Berta et al. 2012; Swain et al. 2012; Deming et al. 2013 ) and Near Infrared Camera and Multi-Object Spectrometer (NICMOS) (Swain et al. 2009a; 2009b; Tinetti et al. 2010; Gibson et al. 2011; Crouzet et al. 2012)....

    [...]

  • ...Water has the largest improvement in uncertainty due to the leverage provided by the WFC3 spectra which covers the 1.15 µ and 1.4 µm water bands....

    [...]

Related Papers (5)