Nothing Special   »   [go: up one dir, main page]

A Sub-Neptune Exoplanet With A Low-Metallicity Methane-Depleted Atmosphere and Mie-Scattering Clouds

Download as pdf or txt
Download as pdf or txt
You are on page 1of 45

A Sub-Neptune Exoplanet with a Low-Metallicity Methane-Depleted Atmosphere

and Mie-Scattering Clouds


BJÖRN BENNEKE1, HEATHER A. KNUTSON2, JOSHUA LOTHRINGER3, IAN J.M. CROSSFIELD4,
JULIANNE I. MOSES5, CAROLINE MORLEY6, LAURA KREIDBERG7, BENJAMIN J. FULTON2,17,
DIANA DRAGOMIR4,18, ANDREW W. HOWARD8, IAN WONG9, JEAN-MICHEL DÉSERT10, PETER R.
MCCULLOUGH11,ELIZA M.-R. KEMPTON12,13,, JONATHAN FORTNEY14, RONALD GILLILAND15,
DRAKE DEMING12, JOSHUA KAMMER16
[1] Department of Physics and Institute for Research on Exoplanets, Université de Montréal, Montréal, QC, Canada
[2] Division of Geological and Planetary Sciences, California Institute of Technology, Pasadena, CA 91125, USA
[3] Lunar & Planetary Laboratory, University of Arizona, 1629 E. University Boulevard., Tucson, AZ, USA
[4] Department of Physics and Kavli Institute of Astronomy, Massachusetts Institute of Technology, 77 Massachusetts Ave,
Cambridge, MA, 02139, USA
[5] Space Science Institute, 4750 Walnut Street, Suite 205, Boulder, CO 80301, USA
[6] Department of Astronomy, University of Texas, Austin, TX 78712, USA
[7] Department of Astronomy, Harvard University, 60 Garden Street, Cambridge, MA 02138, USA
[8] Department of Astronomy, California Institute of Technology, Pasadena, CA 91125, USA
[9] Department of Earth, Atmospheric, and Planetary Sciences, Massachusetts Institute of Technology, 77 Massachusetts
Ave, Cambridge, MA, 02139, USA
[10] Anton Pannekoek Institute for Astronomy, University of Amsterdam, 1090 GE Amsterdam, The Netherlands
[11] Department of Physics and Astronomy, Johns Hopkins University, Baltimore, MD 21218, USA
[12] Department of Astronomy, University of Maryland, College Park, MD 20742, USA
[13] Department of Physics, Grinnell College, 1116 8th Avenue, Grinnell, IA 50112, USA
[14] Department of Astronomy, University of California, Santa Cruz, CA 95064, USA
[15] Space Telescope Science Institute, 3700 San Martin Dr., Baltimore, MD 21218, USA
[16] Southwest Research Institute, San Antonio TX, USA
[17] IPAC-NASA Exoplanet Science Institute Pasadena, CA 91125, USA
[18] NASA Hubble Fellow

With no analogues in the Solar System, the discovery of thousands of exoplanets with masses
and radii intermediate between Earth and Neptune was one of the big surprises of exoplanet
science. These super-Earths and sub-Neptunes likely represent the most common outcome
of planet formation1,2. Mass and radius measurements indicate a diversity in bulk
composition much wider than for gas giants3; however, direct spectroscopic detections of
molecular absorption and constraints on the gas mixing ratios have largely remained limited
to planets more massive than Neptune4–6. Here, we analyze a combined Hubble/Spitzer Space
Telescope dataset of 12 transits and 20 eclipses of the sub-Neptune GJ 3470 b, whose mass of
12.6 M⊕ places it near the half-way point between previously studied exo-Neptunes (22-23
M⊕)5–7 and exoplanets known to have rocky densities (7 M⊕)8. Obtained over many years,
our data set provides a robust detection of water absorption (>5σ) and a thermal emission
detection from the lowest irradiated planet to date. We reveal a low-metallicity, hydrogen-
dominated atmosphere similar to a gas giant, but strongly depleted in methane gas. The low,
near-solar metallicity (O/H=0.2-18) sets important constraints on the potential planet
formation processes at low masses as well as the subsequent accretion of solids. The low
methane abundance indicates that methane is destroyed much more efficiently than
previously predicted, suggesting that the CH4/CO transition curve has to be revisited for
close-in planets. Finally, we also find a sharp drop in the cloud opacity at 2-3 µm
characteristic of Mie scattering, which enables narrow constraints on the cloud particle size
and makes GJ 3470b a keystone target for mid-IR characterization with JWST.
We observed the transiting sub-Neptune mass exoplanet GJ 3470b with the Hubble Space
Telescope (HST) as part of a spectral survey of atmospheres of low mass exoplanets (GO 13665).
With an orbital period of 3.3 days and a mass of 12.6 M⊕, GJ 3470b is a typical member of the
intriguingly abundant class of close-in sub-Neptunes (Figure 1). GJ 3470b’s low surface gravity
combined with the proximity and small size of its host star make it an outstanding candidate for
detailed atmospheric characterization, especially in the sub-Neptune mass regime for which robust
molecular detections have remained elusive to date4,9,10. We obtained time-series spectroscopy of
six transits using HST, including three from 1.1 µm to 1.7 µm using Wide Field Camera 3 (WFC3)
and three at optical wavelengths (0.55–1.0 µm) using the Space Telescope Imaging Spectrograph
(STIS). We complement these HST transit observations with a total six Spitzer/IRAC transit
observations as well as a total of 20 secondary eclipse observations at 3.6 and 4.5 µm (Figure 2
and 3).

We jointly analyze all HST and Spitzer transit data to obtain a consistent visible-to-IR transmission
spectrum covering a wavelength range between 0.55 and 5.0 µm (Figure 2). The details of this
analysis are described in the Methods Section. For each instrument, we verify that the
measurements are consistent over multiple epochs by demonstrating repeatability among the three
transits in each channel, setting tight upper limits on the effect of star spots on the overall spectrum
(Supplementary Figure 4). The transit depth precision with WFC3 is substantially higher than the
one obtained with STIS and Spitzer due to GJ 3470’s higher photon flux in the WFC3 bandpass
and the substantially higher throughput of WFC3 (30-40%) compared to STIS (8-12%). The
achromaticity of the dominant WFC3 systematics further improves the transit depth precision in
the WFC3 spectroscopic channels relative to STIS. Our three transit observations using optimized
WFC3 spatial scans across the full sub-array enable us to collect 30 times more photons and
achieve 5 times higher precision than the previously published stare-mode transit observation of
GJ 3470b with this same instrument9.

Our transmission spectrum for GJ 3470b reveals an attenuated but statistically significant water
absorption feature at 1.4 µm in the WFC3 data, protruding over an otherwise cloud opacity-
dominated visible to near-IR transmission spectrum (Figure 2). The water absorption is detected
in multiple neighboring spectroscopic channels covering the 1.4 µm water band. Quantitively,
retrieval models that include molecular absorption by water are favored by the Bayesian evidence11
at 124,770:1 (5.2σ) and result in significantly better best fits than models without water (see
Methods). Comparisons to models show that the data are best matched by a low-metallicity,
hydrogen-dominated atmosphere (O/H = 0.2–18 x solar) with water vapor absorbing above high-
altitude clouds that become optically thick below the 1 mbar level at 1.5µm (Figure 4). High
metallicity atmospheres with high mean molecular mass are ruled out by the data because the
associated smaller scale height would not allow for the observed transit depth variations. This
finding is independent of the detailed atmospheric models because the observed transit depth
variations would require the cut-off altitude of the grazing star light to vary over greater than 20
atmospheric scale heights across the near-infrared, which is not realistic. The Spitzer eclipse
observations add to the water constraints because substantially increased water opacity would not
allow for the observed contrast in thermal emission at 3.6 and 4.5 µm.
Intriguingly, the high-altitude clouds on GJ 3470b are not well represented by Rayleigh hazes as
previously reported12,13 or a simple gray cloud deck. Instead, our measurements provide direct
observational evidence for the characteristic wavelength dependent extinction of finite-sized Mie
scattering aerosol particles (Figure 2). These aerosol clouds become increasingly transparent at
around 2–3 µm enabling us to constrain their effective particle size to 0.60±0.06 µm in the
uppermost layers of the clouds (Figure 4). This particle size estimate provides a rare direct
observational constraint that can guide the further development and verification of physics-driven
cloud and haze models for exoplanets. Here, we account for the non-gray cloud opacities in our
retrieval analysis by modeling the finite-sized particles using Mie scattering theory11,14 and
describing the effective particle size, the upper cloud deck pressure, and the cloud scale height as
free parameters (see Methods). All posterior constraints on the atmospheric gases provided in this
work account for the uncertainties introduced by the clouds as well as the parameterized “free”
temperature structure (see Supplementary Figures 6 and 7).
GJ 3470b's transmission spectrum also shows a striking absence of methane absorption. For the
relatively cool, low-metallicity atmosphere of GJ 3470b, atmosphere models with solar carbon-to-
oxygen ratio would have predicted methane to be the dominant carbon-bearing molecule.
However, methane absorption at 1.6µm in the WFC3 bandpass and at 3.3µm in Spitzer/IRAC
channel 1 is not observed indicating a strong depleting of methane (Figure 2). The low methane
abundance is independently supported by the twenty secondary eclipse observations, which can be
used to constrain the shape of GJ 3470b's thermal emission spectrum in the 3.6 and 4.5 µm Spitzer
bands (Figure 3). We detect strong thermal flux emerging at 3.6 µm at 4.7σ significance (𝐹# ⁄𝐹∗ =
115,*-
)*+ p.p.m) and a tight upper limit at 4.5 µm (𝐹# ⁄𝐹∗ = 3 ± 22 p.p.m). This is in contrast to the
prediction for the fiducial solar abundance model, but in agreement with the low methane
abundance inferred from the transmission spectrum (Figure 2). Quantitatively, our retrieval
analysis shows that the methane abundance is below 1.3x10-5 at greater than 99.7% confidence,
substantially below the value of 4.6x10-4 expected for a solar abundance atmosphere in chemical
equilibrium (Figure 4). The best fitting models show a striking methane depletion by three orders
of magnitude compared to equilibrium.

We assess the origin of the methane depletion through state-of-the-art photochemical modeling
and thermal modeling of GJ 3470b’s atmosphere (see Methods). Consistent with Refs 15,16 we find
that the methane abundance in the layers probed by the observations should not be reduced
substantially by photochemistry in layers probed by our observations (Figure 4, Supplementary
Figure 8). Possible explanations for the unexpected lack of methane could be substantial interior
heating, photochemical depletion due to catalytic destruction of CH4 in deeper atmospheric
regions, or a low C/O ratio as a result of the planet formation process. The interior heating scenario
would require interior temperatures (Tint) above 300 K to push the otherwise relatively cold mid-
atmosphere of GJ 3470b into the CO dominated regime17,18. Evolution modeling of GJ 3470b
indicates that internal heat from formation should have been radiated away within a few Myr, well
below the estimated age of the system19; however, tidal heating due to forced eccentricity from
another unseen planet in the system, similar to the situation with Jupiter's moon Io could be a
possible explanation. The residual non-zero eccentricity of GJ 3470b as independently confirmed
by our eclipse observations and radial velocity measurements support this hypothesis.
Alternatively, GJ 3470b's surprising lack of methane could potentially be the results of
photochemical depletion due to catalytic destruction of CH4 in deeper atmospheric regions where
photolysis of NH3 and H2S release large amounts of atomic hydrogen. The fact that ammonia is
also depleted in comparison to expectations based on our chemical-kinetics modeling (Figure 4)
is consistent with this catalytic-destruction possibility. Ammonia is an important quenched
disequilibrium product (Supplementary Figure 8), and we would have expected to see NH3
absorption at 1.5 µm (Figure 2). In either scenario, the carbon freed from methane would most
likely be locked up in CO, which can be seen in the transmission spectrum at 4.5µm (Figure 2)
and as a suppression of thermal flux within the 4.5µm Spitzer bandpass (Figure 3). HCN is also
one potential major sink of the carbon in the coupled CH4-NH3 photochemistry if the elemental
N/C ratio is larger than solar20,21; however, we also obtain an upper bound on the HCN abundances
(Supplementary Figure 7). In either case, all CH4 destruction scenarios would also lead to the
production of CH3 and other radicals, some fraction of which can react with other atmospheric
carbon and nitrogen species to form increasingly heavy hydrocarbons and nitriles, eventually
ending up in refractory soot-like haze particles. These photochemically produced particles could
provide an explanation for the observed cloud opacity at shortward of 3 µm, although recent
experimental work has also shown that similar photochemical hazes can be formed even in the
absence of methane22,23. In either case, the 30-90 nm particles found in experiments2 would likely
need to coagulate to form larger aggregates to explain the inferred wavelength dependence of the
cloud opacity on GJ 3470b.

Overall, our spectra show directly through atmospheric observations that close-in sub-Neptunes
can have near-solar metallicity atmospheres likely formed by the direct accretion of primordial gas
from the protoplanetary disk onto a rock/iron or ice-dominated core as suggested by recent planet
formation models19,24. The near-solar metallicity is particularly intriguing because the steep
increase in planet occurrence rate from >20 M⊕ towards 10 M⊕25 suggests that sub-Neptunes
could have a more efficient planet formation process intrinsically different from planets more
massive than Neptune4–7,10,17,26. Sub-Neptune formation beyond the ice line and subsequent
migration could have led to much higher atmospheric metallicities, or even water worlds27, which
we do not find for GJ 3470b. Instead, our measurement of a near-solar water abundance favors
formation scenarios in which the core accreted a primordial gas envelope whose metal content was
subsequently not notably enriched by planetesimal accretion28, the erosion of the icy/rocky core,
or late collisions29. GJ 3470b’s gas envelope is also expected to have undergone substantial mass
loss30, but this mass loss was likely in the hydrodynamic regime as we see no evidence for
preferential loss of lighter elements. Finally, the unexpected methane depletion further indicates
that our understanding of the chemical and thermal processes on these low-mass planets remains
incomplete. Opportunely, our detection of sub-µm Mie scattering particles indicates that the clouds
or hazes in the atmosphere of GJ 3470b become increasingly transparent beyond 3 µm, making
GJ 3470b an excellent target for future JWST mid-IR observations—both as an archetype for the
intriguing population of sub-Neptunes and as a laboratory for atmospheric chemistry and cloud
formation in warm atmospheres.
Acknowledgement. This work is based on observations with the NASA/ESA HST, obtained at
the Space Telescope Science Institute (STScI) operated by AURA, Inc. We received support for
the analyze by NASA through grants under the HST-GO-13665 program (PI Benneke). This work
is also based in part on observations made with the Spitzer Space Telescope, which is operated by
the Jet Propulsion Laboratory, California Institute of Technology under a contract with NASA (PIs
Knutson and Désert). B.B. further acknowledges financial supported by the Natural Sciences and
Engineering Research Council (NSERC) of Canada and the Fond de Recherche Québécois—
Nature et Technologie (FRQNT; Québec). J.M. acknowledges support from NASA grant
NNX16AC64G, the Amsterdam Academic Alliance (AAA) Program, and European Research
Council (ERC) under the programme Exo-Atmos (grant agreement no. 679633). D. D.
acknowledges support provided by NASA through Hubble Fellowship grant HST-HF2-
51372.001-A awarded by the Space Telescope Science Institute, which is operated by the
Association of Universities for Research in Astronomy, Inc., for NASA, under contract NAS5-
26555.

Author contributions. B.B. led the data analysis of the HST and Spitzer transit data, with
contributions from J.L., I.W., and H.K. L.K. and J.M.D. performed independent analyses of the
Spitzer transits and found consistent results. H.K. led the data analysis of the Spitzer secondary
eclipse observations. J.M. provided the chemical kinetics atmosphere models. B.B. and C.M.
provided the self-consistent atmospheric models and the atmospheric retrieval analysis. B.B. wrote
the manuscript, with contributions from B.F., H.K. and J.M. All authors discussed the results and
commented on the draft.

Competing interests. The authors declare no competing financial interests.

Author information. Correspondence and requests for materials should be addressed to B.B.
(bbenneke@astro.umontreal.ca).
Figure 1: Planet mass versus equilibrium temperature for known low-mass planets with signal-to-
noise ratio >3 mass measurements. Planets with published space-based spectroscopic observations for
atmospheric characterization are indicated by large circles (black and coloured) and all other planets by
small grey circles. Equilibrium temperatures are calculated for a Bond albedo of AB = 0.1. Vertical and
horizontal bars indicate the 1σ uncertainties. Among the spectroscopically studied planets, the planets with
non-detections of the atmosphere are shown in black, planets with detection of only transit features in blue,
planets with detected thermal emission deviant from black-body radiation in yellow, and GJ 3470 b with
spectral features detected both in transit and eclipse measurements in red. The green dotted curves show
the equilibrium temperatures for which we expect CH4 (below the curve) and CO (above the curve) to be
the dominant carbon-bearing species in the photosphere based on self- consistent modelling in chemical
equilibrium. The dashed vertical line indicates the mass of Neptune for reference.
Figure 2: Transmission spectrum of GJ 3470b. Black data points show transit depth measurements from
the HST/STIS, HST/WFC3, and Spitzer/IRAC observations analyzed in this study. Vertical and horizontal
black bars indicate the 1σ transit depth uncertainties and the wavelength ranges of the measurements,
respectively. The best fitting model with near-solar water abundance, Mie scattering clouds, and strong
methane depletion is shown by the red curve, with circles indicating the bandpass integrated model. Water
absorption results in increased transit depth at 1.4 µm (zoom in panel b). Finite-sized Mie-scattering
particles (~0.6 µm) result in a characteristic drop off in cloud opacity beyond 2 µm (red dotted curved).
Adding 100 ppm methane to the best fit model results in significant disagreement to the data at 1.6 and 3.6
µm (blue curve). Similarly, adding 100 ppm ammonia results in disagreement at 1.5 µm (green curve in
panel b). A cloud-free solar metallicity model (orange curve) and the best-fitting gray cloud model (gray
dashed curve) are shown in panel (a) for reference. Both provide a poor fit to the data. The dominant
molecular absorbers for each model are labeled at the top with colors matching the color of the spectra. A
distribution of models from the joint retrieval modeling of transit and eclipse data is shown in
Supplementary Figure 6. Previous measurements9,12,31–34 are in statistical agreement with our data, but have
significantly larger transit depth uncertainties and are omitted here for clarity.
self-consistent model
(1xsolar, C/O=0.54)
best-fit to transit only
(CH4-depleted)
best-fit to transit & eclipse
(CH4-depleted)

10
00
K
90
0
K

800
K

700 K

600 K

500 K

Figure 3: Thermal emission spectrum of GJ 3470 b. Spitzer/IRAC measurements (black) are compared
to simulated model atmospheres (colored solid curves) and black body curves (grey dotted curves). Vertical
and horizontal black bars indicate the 1σ transit depth uncertainties and the wavelength ranges of the
measurements, respectively. The horizontal black dashed line represents the 3σ upper limit for the 4.5 µm
measurement. Colored circles indicate the bandpass integrated models. Consistent with the transit data, the
fiducial methane-rich, chemical equilibrium model with solar metallicity (orange) results in a poor fit to the
data. Instead, the strong flux reversal between 3.6 µm and 4.5 µm Spitzer/IRAC bandpasses strongly favors
methane-depleted scenarios. The best joint fit of transit and eclipse with near-solar water abundance and
strong methane depletion matches both data points within 1σ (red, also see red curve in Figure 1).
Consistency between the transit and eclipse is further underscored because the best fit model from the transit
data alone (purple) captures the flux reversal between 3.6 µm and 4.5 µm correctly and presents a much
better predictor of the eclipse data than our fiducial methane-rich self-consistent model (orange).
Transit (a) (b) (c)
Thermal
Transit + Thermal

(d) (e) (f)

Figure 4: Constraints on gas composition and cloud properties in the atmosphere of GJ 3470b. The
six panels show the marginalized probability distributions for the abundances of water vapor (a), methane
(b), ammonia (c), the mean molecular mass (d), the cloud particle size (e), as well as the two-dimensional
marginalized joint probability of the water abundance and the atmospheric pressure at which the clouds
become optically thick (τ=1) (f). The blue 1D distributions indicate the posterior constraints derived from
a joint analysis of all our transit and eclipse data. The blue horizontal bar indicates the central 68%
probability interval. Green and orange show the equivalent distributions based on only the transit and only
the eclipse data, respectively. For comparison, the vertical dashed lines indicate the molecular abundances
for a 1 x solar metallicity atmosphere in chemical equilibrium (red) and based on the full photochemistry
kinetics model (purple). Black lines in panel f indicate the 1σ , 2σ, and 3σ probability contours and the
vertical and horizontal bars illustrate the 1σ uncertainties on the water vapour abundance and cloud top
pressure individually. The water vapor abundance and mean molecular mass is found to be consistent with
a gas giant-like atmosphere with a near-solar composition. Methane is found to be depleted by at least a
factor of 30-50 at greater than 99.7% confidence. Ammonia is found to be depleted at >95% confidence
compared to the expectation from chemical kinetics. Panel f shows the correlation between the water
abundance and clouds due to their opposing effects on the strength of the 1.4µm water band in transit. The
transit data also provide sharp constraints on the cloud particle size (panel e). A corner plot of all
marginalized distributions is shown in Supplementary Figure 7.
Methods
We jointly analyze all twelve HST and Spitzer transits taken between 2012 and 2017
(Supplementary Table 1) in our modular Exoplanet Transits Eclipses & Phasecurves (ExoTEP)
framework. ExoTEP first reduces all raw data into standardized spectroscopic light curves and
then determines the visible-to-IR transmission spectrum by computing the joint posterior
probability of the transit and systematics model parameters of all instruments in one joint MCMC
analysis. The 20 secondary eclipses per Spitzer channel are similarly fit in a global analysis of the
planet’s dayside emission data. Finally, we interpret GJ 3470b’s transmission spectrum and the
broadband Spitzer/IRAC secondary eclipse observations using a free atmospheric retrieval as well
as a suite of self-consistent radiative/convective heat transfer and (dis-)equilibrium chemistry
models. We performed a joint retrieval analysis of all data together as well as transits, eclipses,
and subsets of the transit data individually. To minimize the sensitivity of our results to prior
assumptions on the atmospheric processes, we focused on a “free” retrieval analysis with freely
parameterized molecular abundances, a freely parameterized temperature structure, and freely
parameterized properties of the upper cloud/haze deck.

Observations and data reduction


HST/WFC3 transits.
Each of three HST/WFC3 transit observations consisted of four 96-minute telescope orbits, with
48-min gaps in phase coverage between target visibility periods due to Earth occultation. The
HST/WFC3 data were taken using the G141 grism and the spatial scan mode, which scans the
telescope during the exposure and thereby moves the spectrum perpendicularly to the dispersion
direction on the detector4,35. The spatial scanning enables longer exposures and reduces the
instrumental overhead time compared to staring mode observations. We optimized the
observations efficiency by performing spatial scans of maximum length across almost the full
detector sub-array (256x256). We observed forward and backward scans to further reduce the
instrumental overhead. In this way, we achieved an overall duty cycle of 73% for the bright star
GJ 3470b, collecting more than ten times the number of photons per orbit than the previously
obtained WFC3 stare mode observations of this star with an overall integration efficiency of 6.8%9.
After combining all three transits, we collected 32 times more photons than this previous study,
resulting in a five-fold improvement in the precision of our transit depth measurement in the WFC3
bandpass.

As a result of the spatial scans and the grism dispersion, the star light was spread over a near
rectangular, but slightly trapezoidal patch on the detector. The trapezoidal shape of the illuminated
patch on the detector results from the slight change in dispersion on the detector as the star moves
along the detector’s y-axis, leading to a 2–3-pixel shift in the x direction for a given wavelength.
As in previous studies35, we minimize the sky background contribution in our full-frame images
by co-adding the differences between consecutive up-the-ramp samples only for regions
containing the target spectrum during the time between the consecutive nondestructive reads. We
also flat-fielded the data using the wavelength-dependent flat field data provided by STScI and
replace bad pixels by interpolating spatially.

To account for the trapezoidal shape, we first computed the 2D wavelength solution for spatial
scans on the detector36, 37. We then compute lines of constant wavelength and define trapezoidal
patches based on our predefined wavelength bins. We integrate over each of these trapezoidal
patches for every time step to obtain the uncorrected spectrophotometric light curves. For each
frame, we apply a small correction to the x position of the patch boundaries to account for the
small drift of the star throughout the transit observation. Unlike previous studies10,35,36, we do not
pre-smooth the data in the dispersion direction. Instead, we integrate the flux across the trapezoidal
patch by directly adding the pixel counts for all pixels that are fully within the trapezoidal patch,
including fractions of the pixel counts for pixels that are intersected by the line of constant
wavelength. The exact fraction added to the wavelength bin is determined by performing a local
sub-pixel spline interpolation at the location of the intersected pixels and then adding fractions of
the total counts proportional to the 2D integral of that spline to adjacent wavelength bins. This
procedure ensures that the total flux is conserved. Combined with light curve fitting discussed
below, this method delivers photon-noise limited residuals with no sign of contamination for both
the white light and spectroscopic light curves (Supplementary Figures 1, 11, 12, 13).

HST/STIS transits
We observed three transits of GJ 3470b using the G750L grism on HST/STIS to measure
GJ 3470b’s transmission spectrum between 0.55 and 1.0 µm. The STIS observations were taken
in stare mode, i.e. without moving the telescope. The relatively small overhead for the read-out of
our 1024x128 subarray resulted in a duty cycle of 89.0%. We performed the initial reduction of
the raw data using the latest version of CALSTIS, and removed bad pixels and cosmic ray hits
using a custom-made Python routine that searched for outliers in both position and time. For each
image frame, we then extracted the 1D spectrum using spectral aperture extraction across a 25-
pixel wide aperture. The initial wavelength calibration was provided by CALSTIS. We corrected
for sub-pixel wavelength shifts in the dispersion direction over the course of the visit by finding
the best fitting offset and amplitude between each 1D and the median 1D spectrum of the visit. We
used the amplitudes from this optimization as the data points of the white light curve, as in previous
studies 35. Finally, we used the wavelength-corrected 1D spectra from each exposure to extract the
flux value for 50-nm-wide spectral bandpasses, which we then combined to form the data points
in the photometric light curves for each spectral channel.

Spitzer/IRAC transits and eclipses


We analyzed a total of 6 Spitzer transits and 20 Spitzer eclipses of GJ 3470b in this work: 3 transits
and 10 eclipses in each IRAC channel at 3.6 and 4.5 µm. Two of the 4.5µm Spitzer transits were
previously published38, the remaining 24 visits are previously unpublished. We used the standard
peak-up pointing mode for the observations, which places the star reliably in the center of a pixel
after allowing for an initial 30-minute settling time at the new pointing. We observed our target in
subarray mode with 0.4 s exposures in both bandpasses, with a total duration of 6.44 hours (54,144
images) and 6.45 hours (54,272 images) of science observations for each 3.6 and 4.5 µm visit,
respectively. Following standard procedure, we use the flat-fielded and dark-subtracted “Basic
Calibrated Data” (BCD) images provided by the standard Spitzer pipeline for our analysis and
estimated the sky background following Ref. 39. We then determined the position of the star in
each image using flux-weighted centroiding with a radius of 3.5 pixels, and calculate the flux in a
circular aperture with radii of [2.0, 2.1, 2.2, 2.3, 2.4, 2.5, 2.6, 2.7, 2.8, 2.9, 3.0, 3.5, 4.0, 4.5, 5.0]
pixels to create our photometric time series. We also consider an alternative version of the
photometry utilizing a time-varying aperture, where we scale the radius of the aperture
proportionally to the square root of the noise pixel parameter, which is proportional to the full
width half max (FWHM) of the stellar point spread function 17,40,41. As discussed in Refs. 39,42, we
later optimize our choice of aperture, bin size, and trim duration individually for each visit by
selecting the options which simultaneously minimize the RMS of the unbinned residuals as well
as the time-correlated noise in the data in the individual fits. We quantify this time-correlated noise
component by calculating the RMS as a function of bin size in steps of powers of two points per
bin, and pick the best aperture, bin size, and trim duration that minimizes the least-squares
difference between the actual RMS vs bin size and the white noise prediction for the individual
visit.

Transit white light curve fitting


We perform a global analysis of all WFC3, STIS, and Spitzer transit light curves by simultaneously
fitting our transit light curve model, the instrument systematics models, and photometric noise
parameters within one joint Markov Chain Monte Carlo analysis. Our analysis framework,
ExoTEP, enables us to perform this analysis from extracted raw photometry to the transit
parameters and their uncertainties as a single, statistically consistent Bayesian analysis. The main
outputs of the analysis are the global transit parameters (a/R*, b) as well as the planet-star radius
ratios in the STIS bandpass (0.55-1.01µm), the WFC3 bandpass (1.1-1.7µm), and the two IRAC
bandpasses (3.6 and 4.5 µm). To help the convergence within this high-dimensional parameter
space, ExoTEP first fits each transit light curve individually and then automatically uses the best-
fitting systematics model parameters from those individual fits as initial conditions in the global
MCMC fit. We opt to fit GJ 3470b’s ephemeris (T0, P) with only the Spitzer transit observations
because they cover the transits continuously at high cadence and are spaced over 5 years; hence,
providing exquisite constraints on the GJ 3470b’s transit ephemeris.
HST/WFC3 Instrument Model
Our uncorrected WFC3 transit light curves exhibit well-documented systematic trends in flux with
time, including visit-long slopes and orbit-long exponential ramps 4,35,43. We account for these
systematics by simultaneously fitting the data with the transit model and the analytical model-
ramp function
𝑆3456 (𝑡) = (𝑐𝑑(𝑡) + 𝑣𝑡> ) ⋅ (1 − exp(−𝑎𝑡EFG − 𝑏)) (1)
to correct for these instrumental systematics 26,43. Here, 𝑆3456 (𝑡) is the transit model computed
for the WFC3 bandpass, 𝑐 is a normalization constant, 𝑑 (𝑡) is 1 for forward scans and 𝑑 for
backward scans, 𝑣 is the visit-long linear slope, a and b are the rate constant and amplitude of the
orbit-long exponential slope, and 𝑡> and 𝑡EFG are the time in hours since start of the visit and the
start of the observations within the current orbit. Following standard procedure, we discard the
first HST orbit of each visit in the analysis because the amplitude of the ramp is larger than for the
remaining orbits. We also remove the first forward and first backward scan exposures of each
orbit, which improves the overall light curve fit.
HST/STIS Instrument Model
Our STIS transit light curves exhibit ramp-like systematic trends comparable to those seen in
previous STIS exoplanet data sets. STIS data show a visit-long slope and orbit-long ramps that
result from the temperature settling of the telescope and the reinitiating of the read-out sequence
at the beginning of each orbit. Following standard procedure44–48, we correct for these systematics
by simultaneously fitting the transit model and the analytical systematics model
* 6 N
𝑆IJKI (𝑡) = (𝑐 + 𝑣𝑡> ) ⋅ (1 + 𝑝M 𝑡EFG + 𝑝* 𝑡EFG + 𝑝6 𝑡EFG + 𝑝N 𝑡EFG ) (2)
to the each STIS transit data set. Here, 𝑐 and 𝑣 are again the normalization constant and the visit-
long linear slope, and 𝑝M to 𝑝N are coefficients to describe the systematic trend within each orbit
via a 4th order polynomial function. As in the analysis of WFC3, we discard the first HST orbit of
each visit and remove the first two exposures of each orbit. We explored a wide range of more
complex systematics models that simultaneously detrend against the x and y position and slope of
the spectral trace on the detector as discussed in detail in Ref. 48, but found no substantial
improvement in scatter or changes in our conclusions.
Spitzer/IRAC Instrument Model
Our Spitzer/IRAC instrument model accounts for intra-pixel sensitivity variations and temporal
sensitivity changes using the modified pixel-level decorrelation (PLD) model described in Ref 42.
Our instrument model is
∑^\_M 𝑤\ 𝐷\ (𝑡P ) (3)
𝑆IOPQRSF ( 𝑡P ) = 1 + 𝐴𝑒 )QV /X + 𝑚𝑡 + ,
∑^\_M 𝐷\ (𝑡P )
where the systematics model 𝑆IOPQRSF (𝑡P ) is composed of a fitted linear-exponential ramp in time
(𝐴𝑒 )QV /X + 𝑚𝑡) and the pixel-level decorrelation (PLD) term49. The 𝐷\ (𝑡P )'s in the PLD term are
the raw counts in the 3x3 pixels, k=1…9, covering the central region of the PSF. In the numerator,
these raw data values are multiplied by nine time-independent PLD weights, {𝑤M … 𝑤^ }, fitted as
free parameters in the light curve analysis. Together with the linear slope 𝑚, the instrument model
therefore includes 10 free instrument fitting parameters to capture the intrapixel sensitivity
variations and temporal sensitivity changes.
Transit Model and MCMC Analysis
We compute the transit light curve 𝑓 (𝑡P ) using the Batman implementation50 of the standard transit
equations51. In our joint white light curve fit, we allow for four different transit depths in the
WFC3, STIS, IRAC 3.6µm, and IRAC 4.5µm bandpasses, but jointly fit a single set of transit
geometry parameters (a/R*, b) to be consistent across the entire data set. The limb darkening
coefficients are computed specifically for the STIS, WFC3, and Spitzer/IRAC bandpasses
following the same procedure detailed in Ref. 48 using the PHOENIX stellar models52 for a variety
of different parameterizations53. Our fiducial stellar parameters for GJ 3470 are in T = 3652±50 K,
log(g) = 4.843±0.035, and Z = 0.2±0.1 dex 34,54,55, and we perform the same wide sensitivity tests
as detailed Ref. 48. We account for the cadence of the observations by numerically integrating in
time from the start to end of the individual exposures. We also account for the small but non-zero
eccentricity independently constrained by the eclipse observations discussed below. Given the
transit model and instrument sensitivity described above, we compute the log-likelihood function
for the joint white light curve as
s pq
𝑛k [𝐷k (𝑡P ) − 𝑆k (𝑡P ) ⋅ 𝑓k (𝑡P )]*
ln ℒ = h i−𝑛k ln 𝜎k − ln 2𝜋 − h r , , (4)
2 𝜎k*
k_M P_M

where the log likelihood is summed over all 𝑁 visits including the three WFC3, the three STIS,
the six Spitzer/IRAC transits. The contribution from each visit is computed by summing the
likelihood for each of the 𝑛k data points, 𝐷k (𝑡P ), and simultaneously fitting a free photometric
noise parameter 𝜎k for each visit. We conservatively opt to fit for independent noise parameters
for each visit to allow for visit-to-visit variations in the scatter in the data. The systematics model
𝑆k (𝑡P ) is different for each visit and free parameters are included in the joint fit according to the
systematics models (Equations 1,2, and 3) for the instrument used in the visit. Based on this
likelihood function, we simultaneously compute the best estimates and joint posterior distribution
of all astrophysical and systematics model parameters using the emcee package56, a Python
implementation of the Affine Invariant Markov Chain Monte Carlo (AI-MCMC) Ensemble
sampler57. We assume flat priors on all parameters and initialize the global fit with the best
estimates from MCMC fits to the individual transits. Fitted white light curves and their residuals
for WFC3, STIS, and Spitzer are shown in Supplementary Figure 1, 2, and 3. We also perform
individual fits to each of the transits. The transit depth measurements at each bandpass are highly
repeatable and consistent within their uncertainties (Supplementary Figure 4). A detailed analysis
of all the residuals are provided in the Supplementary Figures 9 – 17.

Constraints on possible star spot effects


None of our twelve HST and Spitzer light curves show any direct evidence for star spot crossing
events. The strongest constraints present the three HST/WFC3 white light curves which, with a
Gaussian residual scatter of approximately 40 ppm per single data point, would have been highly
sensitive to identifying the crossing of star spots along the planet chord 18. In addition, we also find
that all individual transit depth measurements taken over multiple epochs are consistent within
their statistical uncertainties and with the joint fit (Supplementary Figure 4). Again, HST/WFC3
provides the strongest constraints and shows full consistency of the transit depths within 20 ppm.
This consistency in time provides strong confidence that temporal variations of the star have no
effect on the main conclusion derived in this work. Still, in what follows, we quantify the
constraints using stellar models, and we account for possible common-mode effects in the retrieval
analysis.
To calculate the potential effect of star spots on our results, we conservatively adopt a large spot
coverage of 12.5-percent of the stellar disk and a spot temperature 250 K less than the photosphere,
based on the values as reported for GJ 3470 in Ref. 13. We then represent the stellar atmosphere
inside and outside the star spots by two different Phoenix model atmospheres and allow for the
possibility that some of the spots are occulted by the planet during transit and some are unocculted.
Using this model, we investigate the potential effects of star spots by defining the fractional area
of spots occulted during transit, 𝑓, and exploring the effects on the transmission spectrum across
the whole range of 𝑓 values consistent with the observed repeatability of the WFC3 white light
curve transit depth measurements (Supplementary Figure 4). For example, given the spot coverage
of 12.5-percent, 𝑓 =0.1 means that the spots occulted during transit cover 1.25-percent of the stellar
disk and the unocculted spots cover 11.25-percent of the disk.
Supplementary Figure 4e shows the effect on the planet's inferred transmission spectrum for five
representative values of f . For large values of f, a larger fraction of star spots is occulted during
transit. As a result, less stellar flux is blocked by the transiting planet and the planet radius appears
smaller, especially at the shortest wavelength (green and red lines). The opposite effect occurs for
small values of f; unocculted spots cause the planet to appear larger, especially at the shortest
wavelengths, because an overproportioned fraction of the stellar flux is blocked during transit (blue
and purple curves). Finally, the effects become near zero (black curve) if the fraction of the spots
that are occulted is equal to the fraction of the stellar disk crossed by the planet during the entire
transit (unbiased crossing).
Quantitatively, we assess the effects of this departures from unbiased spot crossing event on the
measured white light curve transit by integrating the star spectrum over the bandpass of the WFC3
instrument. We find that, for fractions f=0.06 and f=0.13, the star spot effects reach as high as 20
ppm in the WFC3 white light curve band pass, the maximum value to be consistent with the
repeatability of the WFC3 white light curve measurements (Supplementary Figure 4a). In these
limiting cases, the departures from unbiased spot crossing by the planet can produce effects on the
planet's spectrum of maximum 150 ppm at STIS wavelengths, but no more than ±15 ppm in the
amplitude of the WFC3 water absorption, and a few ppm at Spitzer wavelengths (Supplementary
Figure 4e). However, even in these extreme cases, the effects of star spots on the WFC3 and Spitzer
points are small compared to the uncertainties of the spectral transit depth measurement and the
strength of the water detection. In addition, we find a consistent three transit measurements across
our three STIS visits, and we deem it to be highly unlikely that a discordant STIS spectrum of the
planet was fortuitously combined with a star spot effect to produce the consistent transit radius
that we observe. We still include these common-mode uncertainties as nuisance parameters in the
retrieval as discussed below.

Transit spectroscopy
The ramp-like instrument systematics in our WFC3 observations are, to first order, independent of
wavelength and position on the detector. Following standard practice, we therefore remove the
ramp-like systematics from the photometric time series of each spectroscopic channel by applying
corrections based on the white light systematics model. We apply this pre-correction to the
spectroscopic time series in two different ways and find negligible differences in the resulting
transit depth fits. The first approach is to divide the spectroscopic time series by the systematics
models given in Eq. (1) using the best-fitting parameters from the white light curve fit 4. The
second approach is to divide the spectroscopic time series by the ratio of the uncorrected white
light time series and the best fitting white-light transit model. The latter approach mathematically
resembles the approach of fitting white-light-divided time series taken in Ref. 35. The advantage
of both approaches is that the spectroscopic time series are pre-corrected using a systematics model
derived from high SNR broadband light curves allowing for fewer systematics parameters to be fit
to spectroscopic light curves. Finally, we obtain the transit depth for each of the spectroscopic
channels by performing joint MCMC fits to the pre-corrected spectroscopic times series from all
three WFC3 transits for that spectroscopic channel. Similar to the white light curve fit, we
simultaneously fit a transit model and a systematics model using MCMC. The only differences are
that the parameters (a/R*, b, T0, P) are fixed to the best-fitting values from the white light curve fit
and that systematics model can be highly simplified because the dominant time-dependent ramp
have already been removed.

We obtain the STIS transmission spectrum by performing independent fits to the times series of
eight 50-nm-wide spectroscopic channels between 0.55 and 0.95 µm. As with WFC3, we fit the
transit depth for each of the spectroscopic channels by performing a joint MCMC fit to all three
STIS transits. Following standard practice, however, we use the full STIS systematics model (Eq.
2) for each of our channels 46,58. We explored pre-correcting the spectroscopic light curves with
the white light curve, but found that the systematics are not sufficiently uniform across the
spectroscopic channels to benefit from this approach. The final transmission spectrum is shown in
Figure 2 and the light curve fit for a typical spectroscopic channel are shown in Supplementary
Figure 2. Within the uncertainties, all STIS transit depths are consistent with a single value and
with the WFC3 data, supporting the conclusion from WFC3 that high altitude clouds are present
in the atmosphere along the terminator. We find no evidence for the Rayleigh scattering slope
reported by 12 within the wavelength range covered by our observations (> 0.55 µm). The transit
depth uncertainties in the STIS are larger than the ones derived from the WFC3 data because the
instrument throughput of STIS is three times lower and the M2 star GJ3470b is approximately 40
fainter in the STIS bandpass than in the WFC3 bandpass.

Secondary eclipse analysis


As for the transits, we fit each Spitzer eclipse time series with the pixel-level decorrelation (PLD)
model42,49. We include both a linear and (for the 3.6 µm data only) an exponential function of time.
Using fits to each individual eclipse, we first identify the optimal choice of aperture, bin size, and
trim duration which minimizes simultaneously minimize the RMS of the unbinned residuals as
well as the time-correlated noise in the data (Supplementary Table 2).
As expected in the Spitzer program design, we do not detect the eclipse with greater than 3s
significance in any individual visit, and therefore carry out a simultaneous fit to the ten eclipse
light curves in each bandpass. We detect the eclipse at 3.6 micron with 4.7 sigma significance and
place an upper limit on the 4.5 micron eclipse depth. The constraints on the eclipse depths and
brightness temperatures are 𝐹# ⁄𝐹∗ = 115,*- ,66
)*+ and 𝑇w = 844)6^ at 3.6 µm and 𝐹# ⁄𝐹∗ = 3 ± 22 and
𝑇w < 532 K at 4.5 µm, respectively. For the global fit we allow all of the instrumental noise model
parameters to vary independently for each visit, but assume a common depth and orbital phase for
the secondary eclipse signal. The eclipse is slightly offset relative to the estimated eclipse time for
a perfectly circular orbit. This is consistent with the radial velocity of GJ 3470b, which
independently confirm a small, but non-zero eccentricity 55. The resulting eclipse depths for the
global and individual eclipse fits are shown in Supplementary Figure 5. Here, we fixed the eclipse
time to the value of the global fit at 3.6µm. Consistent with the uncertainties, the ten individual
measurements are randomly distributed around the global fit. Six of ten and seven of ten data
points are within the 68% confidence interval for the observations at 3.6 and 4.5 µm, respectively.
We note that there are either eleven (4.5 µm) or thirteen (3.6 µm) free parameters for the
instrumental noise model in the individual eclipse fits, including nine pixel-level light curve
coefficients, a linear function of time, the measurement error, and a timescale and a normalization
term for the exponential function (3.6 µm data only). This results in a prohibitively large global
model, which includes 132 free parameters in the case of the 3.6 µm data. We instead elect to
reduce the degrees of freedom in our model by using linear regression to determine the nine best-
fit pixel-level light curve coefficients at each step in our MCMC fit. This reduces the total number
of free parameters to 42 and 22 for the 3.6 and 4.5 µm fits, respectively. While this approach is
not formally correct, as we are effectively optimizing a subset of our model parameters rather than
marginalizing (i.e., integrating the posterior probability distribution over all possible parameter
values), we find that when comparing fits to individual eclipse observations using linear regression
to those utilizing the full PLD model the reported uncertainty on the eclipse depth decreases by
less than 10%.

Atmospheric retrieval
We interpret GJ 3470b’s transmission and thermal spectra using a suite of established modeling
tools, including atmospheric retrieval and thermo- and photochemical kinetics and transport
modeling. As we combined a comprehensive data set of transit and eclipse data covering 0.55 to
5 µm, our analysis provides substantially narrower constraints on the composition than previously
reported12,13,59,60. On the atmospheric retrieval side, we use a derivative of the SCARLET
atmospheric retrieval suite4,5,11,61–63 to interpret the transit and eclipse spectra jointly and
individually. The SCARLET forward model computes star light transmitted through the
atmosphere as well as the upwelling disk-integrated emission given a set of molecular abundances,
the temperature structure, and cloud properties. The forward model is then coupled to the Affine
Invariant Markov chain Monte Carlo (MCMC) Ensemble sampler emcee56 to perform the
estimation of the gas composition, thermal, and cloud parameters. We retrieve the joint posterior
probability distribution of the volume mixing ratios of 6 molecular gases (H2O, CH4, CO, CO2,
NH3, and HCN), 5 parameters for the temperature structure64,65, 3 parameters for Mie-scattering
cloud particles, and 2 nuisance parameters for common-mode uncertainties within the transit
depths measurements from WFC3 and STIS. Our Mie scattering cloud parameterization is newly
developed while our parameterization of the temperature structure and the molecular abundances
are common practice.
Mie scattering cloud parameterization. Since the data show direct evidence for scattering by
finite-sized particles, we developed a new Mie-scattering cloud parameterization for atmospheric
retrieval that enables us to explore the properties of the cloud particles near the top of the clouds.
Mie scattering had previously been widely used in atmospheric forward models of
exoplanets11,14,17,66, but not directly within a Bayesian atmospheric retrieval analysis. Driven by
the observational geometry, the objective of the new parameterization is to capture the wavelength
dependent line-of-sight extinction of a wide range of cloud tops, while remaining low complexity
and independent of preconceived ideas of cloud formation in this previously unexplored planet
regime. We simultaneously accomplish these objectives by describing the clouds using three free
parameters: the effective particle size near the cloud top (𝑅}~•€ ), the scale height or ‘fuzziness’ of
the clouds near the cloud top in units of the local gas density scale height (𝐻}~•€ /𝐻‚~ƒ ), and the
atmospheric pressure at which the clouds become optically thick to grazing light beams at 1.5µm
(𝑃X_M ). For each set of the three cloud parameters, SCARLET then iteratively derives an altitude-
dependent particle number density profile 𝑛}~•€ (𝑧) near the cloud top to match the value for 𝑃X_M ,
the pressure at which the optical depth for grazing light beams reaches 𝜏 = 1. The particle size
distribution is a log-normal particle size distribution centered around 𝑅}~•€ in each layer, and a full
wavelength-dependent Mie scattering computation is performed for each particle size in the log-
normal distribution to model the radiative effects of the finite-sized particles. This process is
repeated for each parameter set suggested by the MCMC. For the wavelength-dependent refractive
index of the clouds material, we assume the ones of KCl, and we find that our results are not
substantial changed when replacing KCl by ZnS, Na2S or soots, which are also possible
compositions for clouds on GJ 3470b67,68 The described three parameter were chosen because they
describe the line-of-sight extinction properties of the cloud top in a highly orthogonal way, ideal
for the retrieval. The parameter 𝑃X_M predominately sets the vertical offset between the cloud
extinction and the molecular absorption in the transmission spectrum, 𝑅}~•€ sets the wavelength-
dependency of cloud opacities, and 𝐻}~•€ /𝐻‚~ƒ predominately sets the amplitude of the extinction
contrast across the transmission spectrum. We deliberately leave out parameters for the cloud base
and lower clouds because they would not be constrained by the data. In the limit of large cloud
particles, the models returns to the assumptions of gray clouds as applied in previous retrieval
analyses4,11,26,69. Similarly in the limit of small particles, the new parameterization returns to the
assumption of Rayleigh scattering hazes (𝜎 ∝ 𝜆)N ).
Parameterization of temperature profile. The temperature–pressure profile is parameterized
with five free parameters using the analytic approximation from Refs. 64,65. The free
parameterization is flexible enough to permit a wide range of thermal structures including thermal
inversions, while resulting in smooth, physically motivated profiles consistent with radiative
energy balance as the analytic approximation enforces radiative equilibrium in its derivation. In
the three retrievals depicted in Figure 4, we allowed for fully independent temperature profiles
near the terminator (transit; green) and on the dayside (secondary eclipse; orange). Only for the
joint fit (blue), we assumed that the temperatures are approximatively similar. We do this because
additional five parameters for the terminator temperature structure is computationally extremely
expensive and not justified given that our transmission spectrum contains virtually no information
on the temperature structure11 and the temperature structure near the terminator is not the main
focus of this work. Hence, the retrieved temperature structure and its uncertainties depicted
Supplementary Figure 6 are dominated by the information in the secondary eclipse measurements.
The increase in temperature from 0.01 bars to 10 bars is constrained by the observed brightness
temperature contrast between the 3.6µm and 4.5µm Spitzer band passes. Still, we additionally
verify that the assumption of similar temperature structure does not significantly affect the
conclusions from our joint retrieval by running independent retrievals based on only the
transmission spectrum and based on transit and eclipse, where we deliberately offset the
temperatures by ±20%. No substantial changes in the retrieval results are observed because free
retrievals that directly specify molecular abundances rather than modeling the chemistry are not
very sensitive to the exact temperature structure 11.
Parameterization of molecular composition. Simultaneous with the cloud and temperature
parameters, we retrieve six thermochemically plausible gases that absorb over the wavelengths
covered by the observations (H2O, CH4, CO, CO2, NH3, and HCN). We assign uniform-in-log
mixing ratio priors spanning from −10 to 0. Molecular hydrogen and helium (in solar proportions)
are assumed to comprise the remaining gas such that all species sum to unity. We use the
absorption cross-section from ExoMol for CH4, NH3, and HCN and the empirical opacities list
from HITEMP for H2O, CO, and CO2 70,71. In the free retrieval, we make no a priori assumptions
on the chemistry of our retrieved abundances.
Nuisance parameters for common-mode uncertainties. The analysis of the spectral light curves
from WFC3 uses common mode corrections derived from the white light curve analysis to increase
the relative precision and reduce the complexity of the systematics models (see above). We account
for the small common-mode offsets that this can introduce between the WFC3 points and the other
instruments by including a free offset nuisance parameter in the retrieval. Similarly, we introduce
a white light curve offset nuisance parameter for STIS to account for the uncertainty introduced
by the possible presence of unocculted star spots. The variance for the WFC3 offset prior is the
WFC3 white light curve uncertainty (5ppm) in square sum of with the constraints from the star
spot assessment (see star spot section). For STIS we only account for the stellar effects because no
common mode correction is used in the analysis.
Significance of molecule detection. We determine the significance of the water detection based
on full Bayesian evidence calculations as developed in Ref 11 and now widely applied, e.g. Refs.
26,72–74
. A molecule is deemed "detected" if the improvement in likelihoods from the addition of
that given parameter outweighs the increase in prior volume. This is the most direct Bayesian
approach to assess the significance of a detection, hence removing the need for the approximations
in the commonly used Bayesian information criterion (BIC) as an estimator for significance. Here,
we determine the significance of the water detection by removing water from the model and
rerunning the retrieval to compute the new Bayesian evidence without water. We find the evidence
ratio of the models with and without water, i.e. the Bayes factor, to be BH2O=124,700
corresponding to a highly robust detection. For reference, a 5.2σ significance would be equivalent
in the frequentist’s framework. The lower Bayesian evidence for the model without water (one
*
parameter less) is a direct result of the difference in maximum likelihood (Δ𝜒‹Œ• = 23.8) as the
model without water cannot match the WFC3 data.

Self-consistent atmosphere models and photochemistry calculations


To put derived retrieval results and observations in context, we use self-consistent forward
modeling tools to derive the expected molecular abundances and spectra for planetary atmospheres
in thermal and chemical equilibrium (Figure 2, 3 and 4). A set of fiducial atmospheres was
computed using the implementation described in Refs. 11,63. Both models iteratively solve the
radiative-convective heat transport and chemical equilibrium and result in virtually identical
abundances and spectra. To investigate the deficiency in methane and ammonia, we then also apply
the thermo- and photochemical kinetics and transport model16,75,76 to assess the effects of non-
equilibrium chemistry. The kinetics and transport model captures the three main chemical
processes in planetary atmospheres: thermochemical equilibrium in the deep atmosphere,
transport-induced quenching in the mid-atmosphere, and photochemistry in the upper atmosphere.
A state-of-the-art reaction list of 1760 chemical reactions for 92 molecular species formed by the
elements H, C, O, and N is adopted16. The temperature structure for the kinetics and transport
simulations is set to the ones derived in the self-consistent forward models. The results from both
the chemical equilibrium and the chemical kinetics/photochemistry calculations are indicated in
Figure 3, demonstrating that the lack of methane (CH4) cannot be explained by currently captured
non-equilibrium effects in our 1D models. Photochemical destruction is expected to result in the
reduction of the methane abundance in the uppermost atmosphere of GJ 3470b (p < 10 µbar), but
the pressure levels that dominate the transmission spectrum are virtually unaffected by the methane
destruction—as previously also shown in a theoretical modeling investigation of the sub-Neptune
GJ 1214b15 and GJ 436b16,21 (Supplementary Figure 8).

Data Availability. The data presented in this work are publicly available in the Mikulski
Archive for Space Telescope (MAST) and the Spitzer Heritage Archive (SHA).
1. Dressing, C. D. & Charbonneau, D. THE OCCURRENCE OF POTENTIALLY HABITABLE
PLANETS ORBITING M DWARFS ESTIMATED FROM THE FULLKEPLERDATASET
AND AN EMPIRICAL MEASUREMENT OF THE DETECTION SENSITIVITY.
Astrophys. J. 807, 45 (2015).
2. Fulton, B. J. & Petigura, E. A. The California-Kepler Survey VII. Precise planet radii leveraging
Gaia DR2 reveal the stellar mass dependence of the planet radius gap. Astron. J. 156, 264
(2018).
3. Fulton, B. J. et al. The California-Kepler Survey. III. A gap in the radius distribution of small
planets. Astron. J. 154, 109 (2017).
4. Kreidberg, L. et al. Clouds in the atmosphere of the super-Earth exoplanet GJ 1214b. Nature
505, 69–72 (2014).
5. Fraine, J. et al. Water vapour absorption in the clear atmosphere of a Neptune-sized exoplanet.
Nature 513, 526–529 (2014).
6. Wakeford, H. R. et al. HAT-P-26b: A Neptune-mass exoplanet with a well-constrained heavy
element abundance. Science 356, 628–631 (2017).
7. Stassun, K. G., Collins, K. A. & Gaudi, B. S. Accurate Empirical Radii and Masses of Planets
and Their Host Stars with Gaia Parallaxes. Astron. J. 153, 136 (2017).
8. Ment, K. et al. A second planet with an Earth-like composition orbiting the nearby M dwarf
LHS 1140. Astron. J. 157, 32 (2019).
9. Ehrenreich, D. et al. Near-infrared transmission spectrum of the warm-Uranus GJ 3470b with
the Wide Field Camera-3 on the Hubble Space Telescope. Astron. Astrophys. 570, A89 (2014).
10. Knutson, H. A. et al. Hubble Space Telescope Near-IR Transmission Spectroscopy of the
Super-Earth HD 97658b. Astrophys. J. 794, 155 (2014).
11. Benneke, B. & Seager, S. How to Distinguish between Cloudy Mini-Neptunes and
Water/Volatile-dominated Super-Earths. Astrophys. J. 778, 153 (2013).
12. Nascimbeni, V. et al. The blue sky of GJ3470b: the atmosphere of a low-mass planet unveiled
by ground-based photometry. Astron. Astrophys. 559, 32 (2013).
13. Chen, G. et al. The GTC exoplanet transit spectroscopy survey - V. A spectrally-resolved
Rayleigh scattering slope in GJ 3470b. Astron. Astrophys. 600, A138 (2017).
14. Wakeford, H. R. et al. High-temperature condensate clouds in super-hot Jupiter atmospheres.
Mon. Not. R. Astron. Soc. 464, 4247–4254 (2017).
15. Miller-Ricci Kempton, E., Zahnle, K. & Fortney, J. J. THE ATMOSPHERIC CHEMISTRY
OF GJ 1214b: PHOTOCHEMISTRY AND CLOUDS. Astrophys. J. 745, 3 (2012).
16. Moses, J. I. et al. Compositional Diversity in the Atmospheres of Hot Neptunes, with
Application to GJ 436b. Astrophys. J. 777, 34 (2013).
17. Morley, C. V. et al. Forward and Inverse Modeling of the Emission and Transmission
Spectrum of GJ 436b: Investigating Metal Enrichment, Tidal Heating, and Clouds. Astron. J.
153, 86 (2017).
18. Kreidberg, L., Line, M. R., Thorngren, D., Morley, C. V. & Stevenson, K. B. Water, High-
altitude Condensates, and Possible Methane Depletion in the Atmosphere of the Warm Super-
Neptune WASP-107b. Astrophys. J. Lett. 858, L6 (2018).
19. Lee, E. J. & Chiang, E. To Cool is to Accrete: Analytic Scalings for Nebular Accretion of
Planetary Atmospheres. Astrophys. J. 811, 41 (2015).
20. Moses, J. I., Madhusudhan, N., Visscher, C. & Freedman, R. S. Chemical Consequences of
the C/O Ratio on Hot Jupiters: Examples from WASP-12b, CoRoT-2b, XO-1b, and HD
189733b. Astrophys. J. 763, 25 (2013).
21. Line, M. R., Vasisht, G., Chen, P., Angerhausen, D. & Yung, Y. L. Thermochemical and
Photochemical Kinetics in Cooler Hydrogen-dominated Extrasolar Planets: A Methane-poor
GJ436b? Astrophys. J. 738, 32 (2011).
22. Hörst, S. M. et al. Haze production rates in super-Earth and mini-Neptune atmosphere
experiments. Nat. Astron. 2, 303–306 (2018).
23. He, C. et al. Laboratory Simulations of Haze Formation in the Atmospheres of Super-Earths
and Mini-Neptunes: Particle Color and Size Distribution. Astrophys. J. 856, L3 (2018).
24. Ginzburg, S., Schlichting, H. E. & Sari, R. SUPER-EARTH ATMOSPHERES: SELF-
CONSISTENT GAS ACCRETION AND RETENTION. Astrophys. J. 825, 29 (2016).
25. Mayor, M. et al. The HARPS search for southern extra-solar planets XXXIV. Occurrence,
mass distribution and orbital properties of super-Earths and Neptune-mass planets. ArXiv E-
Prints 1109, arXiv:1109.2497 (2011).
26. Kreidberg, L. et al. A DETECTION OF WATER IN THE TRANSMISSION SPECTRUM
OF THE HOT JUPITER WASP-12b AND IMPLICATIONS FOR ITS ATMOSPHERIC
COMPOSITION. Astrophys. J. 814, 66 (2015).
27. Rogers, L. A. & Seager, S. THREE POSSIBLE ORIGINS FOR THE GAS LAYER ON GJ
1214b. Astrophys. J. 716, 1208–1216 (2010).
28. Mousis, O. et al. DETERMINATION OF THE MINIMUM MASSES OF HEAVY
ELEMENTS IN THE ENVELOPES OF JUPITER AND SATURN. Astrophys. J. 696, 1348–
1354 (2009).
29. Inamdar, N. K. & Schlichting, H. E. STEALING THE GAS: GIANT IMPACTS AND THE
LARGE DIVERSITY IN EXOPLANET DENSITIES. Astrophys. J. 817, L13 (2016).
30. Bourrier, V. et al. Hubble PanCET: an extended upper atmosphere of neutral hydrogen around
the warm Neptune GJ 3470b. Astron. Astrophys. 620, A147 (2018).
31. Bonfils, X. et al. A hot Uranus transiting the nearby M dwarf GJ 3470. Detected with HARPS
velocimetry. Captured in transit with TRAPPIST photometry. Astron. Astrophys. 546, A27
(2012).
32. Crossfield, I. J. M., Barman, T., Hansen, B. M. S. & Howard, A. W. Warm ice giant GJ 3470b:
I. A flat transmission spectrum indicates a hazy, low-methane, and/or metal-rich atmosphere⋆.
Astron. Astrophys. 559, A33 (2013).
33. Dragomir, D. et al. Rayleigh Scattering in the Atmosphere of the Warm Exo-Neptune GJ
3470b. Astrophys. J. 814, 102 (2015).
34. Biddle, L. I. et al. Warm ice giant GJ 3470b – II. Revised planetary and stellar parameters
from optical to near-infrared transit photometry. Mon. Not. R. Astron. Soc. 443, 1810–1820
(2014).
35. Deming, D. et al. Infrared Transmission Spectroscopy of the Exoplanets HD 209458b and XO-
1b Using the Wide Field Camera-3 on the Hubble Space Telescope. Astrophys. J. 774, 95
(2013).
36. Tsiaras, A. et al. A new approach to analysing HST spatial scans: the transmission spectrum
of HD 209458b. (2015).
37. Tsiaras, A. et al. A Population Study of Gaseous Exoplanets. Astron. J. 155, 156 (2018).
38. Demory, B.-O. et al. Spitzer Observations of GJ 3470 b: A Very Low-density Neptune-size
Planet Orbiting a Metal-rich M Dwarf. Astrophys. J. 768, 154 (2013).
39. Kammer, J. A. et al. Spitzer Secondary Eclipse Observations of Five Cool Gas Giant Planets
and Empirical Trends in Cool Planet Emission Spectra. Astrophys. J. 810, 118 (2015).
40. Knutson, H. A. et al. 3.6 and 4.5 µm Phase Curves and Evidence for Non-equilibrium
Chemistry in the Atmosphere of Extrasolar Planet HD 189733b. Astrophys. J. 754, 22 (2012).
41. Lewis, N. K. et al. Orbital Phase Variations of the Eccentric Giant Planet HAT-P-2b.
Astrophys. J. 766, 95 (2013).
42. Benneke, B. et al. Spitzer Observations Confirm and Rescue the Habitable-zone Super-Earth
K2-18b for Future Characterization. Astrophys. J. 834, 187 (2017).
43. Berta, Z. K. et al. THE FLAT TRANSMISSION SPECTRUM OF THE SUPER-EARTH
GJ1214b FROM WIDE FIELD CAMERA 3 ON THE HUBBLE SPACE TELESCOPE.
Astrophys. J. 747, 35 (2012).
44. Sing, D. K. et al. Determining atmospheric conditions at the terminator of the hot-Jupiter
HD209458b . Astrophys. J. 686, 667 (2008).
45. Nikolov, N. et al. Hubble Space Telescope hot Jupiter transmission spectral survey: a detection
of Na and strong optical absorption in HAT-P-1b. Mon. Not. R. Astron. Soc. 437, 46–66
(2014).
46. Sing, D. K. et al. A continuum from clear to cloudy hot-Jupiter exoplanets without primordial
water depletion. Nature 529, 59–62 (2016).
47. Wakeford, H. R. et al. HAT-P-26b: A Neptune-mass exoplanet with a well-constrained heavy
element abundance. Science 356, 628–631 (2017).
48. Lothringer, J. D. et al. An HST /STIS Optical Transmission Spectrum of Warm Neptune GJ
436b. Astron. J. 155, 66 (2018).
49. Deming, D. et al. SPITZER SECONDARY ECLIPSES OF THE DENSE, MODESTLY-
IRRADIATED, GIANT EXOPLANET HAT-P- USING PIXEL-LEVEL
DECORRELATION. Astrophys. J. 805, 132 (2015).
50. Kreidberg, L. batman: BAsic Transit Model cAlculatioN in Python. Publ. Astron. Soc. Pac.
127, 1161 (2015).
51. Mandel, K. & Agol, E. Analytic Light Curves for Planetary Transit Searches. Astrophys. J.
Lett. 580, L171 (2002).
52. Husser, T.-O. et al. A new extensive library of PHOENIX stellar atmospheres and synthetic
spectra. Astron. Astrophys. 553, A6 (2013).
53. Parviainen, H. & Aigrain, S. ldtk: Limb Darkening Toolkit. Mon. Not. R. Astron. Soc. 453,
3821–3826 (2015).
54. Demory, B.-O. et al. Spitzer Observations of GJ 3470 b: A Very Low-density Neptune-size
Planet Orbiting a Metal-rich M Dwarf. Astrophys. J. 768, 154 (2013).
55. Kosiarek, M. R. Bright opportunities for atmospheric characterization of small planets: masses
and radii of K2-3 b, c, d and GJ3470 b from radial velocity measurements and Spitzer transits.
Astron. J. 157, 97 (2019).
56. Foreman-Mackey, D., Hogg, D. W., Lang, D. & Goodman, J. emcee: The MCMC Hammer.
Publ. Astron. Soc. Pac. 125, 306–312 (2013).
57. Goodman, J. & Weare, J. Ensemble samplers with affine invariance. Commun. Appl. Math.
Comput. Sci. 5, 65–80 (2010).
58. Sing, D. K., Vidal-Madjar, A., Désert, J. -M., Lecavelier des Etangs, A. & Ballester, G. Hubble
Space Telescope STIS Optical Transit Transmission Spectra of the Hot Jupiter HD
209458b. Astrophys. J. 686, 658–666 (2008).
59. Dragomir, D. et al. Rayleigh Scattering in the Atmosphere of the Warm Exo-Neptune GJ
3470b. Astrophys. J. 814, 102 (2015).
60. Fisher, C. & Heng, K. Retrieval analysis of 38 WFC3 transmission spectra and resolution of
the normalization degeneracy. Mon. Not. R. Astron. Soc. 481, 4698–4727 (2018).
61. Benneke, B. & Seager, S. ATMOSPHERIC RETRIEVAL FOR SUPER-EARTHS:
UNIQUELY CONSTRAINING THE ATMOSPHERIC COMPOSITION WITH
TRANSMISSION SPECTROSCOPY. Astrophys. J. 753, 100 (2012).
62. Knutson, H. A., Benneke, B., Deming, D. & Homeier, D. A featureless transmission spectrum
for the Neptune-mass exoplanet GJ 436b. Nature 505, 66–68 (2014).
63. Benneke, B. Strict Upper Limits on the Carbon-to-Oxygen Ratios of Eight Hot Jupiters from
Self-Consistent Atmospheric Retrieval. ArXiv150407655 Astro-Ph (2015).
64. Parmentier, V. & Guillot, T. A non-grey analytical model for irradiated atmospheres: I.
Derivation. Astron. Astrophys. 562, A133 (2014).
65. Line, M. R. et al. A Systematic Retrieval Analysis of Secondary Eclipse Spectra. I. A
Comparison of Atmospheric Retrieval Techniques. Astrophys. J. 775, 137 (2013).
66. Gao, P. & Benneke, B. Microphysics of KCl and ZnS Clouds on GJ 1214 b. Astrophys. J. 863,
165 (2018).
67. Morley, C. V. et al. Quantitatively Assessing the Role of Clouds in the Transmission Spectrum
of GJ 1214b. Astrophys. J. 775, 33 (2013).
68. Morley, C. V. et al. Thermal Emission and Reflected Light Spectra of Super Earths with Flat
Transmission Spectra. Astrophys. J. 815, 110 (2015).
69. Line, M. R., Knutson, H., Deming, D., Wilkins, A. & Desert, J.-M. A Near-infrared
Transmission Spectrum for the Warm Saturn HAT-P-12b. Astrophys. J. 778, 183 (2013).
70. Tennyson, J. & Yurchenko, S. N. ExoMol: molecular line lists for exoplanet and other
atmospheres. Mon. Not. R. Astron. Soc. 425, 21–33 (2012).
71. Rothman, L. S. et al. HITEMP, the high-temperature molecular spectroscopic database. J.
Quant. Spectrosc. Radiat. Transf. 111, 2139–2150 (2010).
72. Swain, M. R., Line, M. R. & Deroo, P. ON THE DETECTION OF MOLECULES IN THE
ATMOSPHERE OF HD 189733b USINGHSTNICMOS TRANSMISSION
SPECTROSCOPY. Astrophys. J. 784, 133 (2014).
73. Waldmann, I. P. et al. Tau-REx II: Retrieval of emission spectra.. Astrophys. J. 813, 13 (2015).
74. Line, M. R. et al. No Thermal Inversion and a Solar Water Abundance for the Hot Jupiter HD
209458b from HST/WFC3 Spectroscopy. Astron. J. 152, 203 (2016).
75. Moses, J. I. et al. Disequilibrium Carbon, Oxygen, and Nitrogen Chemistry in the Atmospheres
of HD 189733b and HD 209458b. Astrophys. J. 737, 15 (2011).
76. Moses, J. I. et al. On the composition of young, directly imaged giant planets. Astrophys. J.
829, 66 (2016).
Supplementary Information

A Sub-Neptune Exoplanet with a Low-Metallicity Methane-Depleted Atmosphere


and Mie-Scattering Clouds
BJÖRN BENNEKE1, HEATHER A. KNUTSON2, JOSHUA LOTHRINGER3, IAN CROSSFIELD4, JULIANNE
MOSES5, CAROLINE MORLEY6, LAURA KREIDBERG7, BJ FULTON2, DIANA DRAGOMIR4,17, ANDREW
HOWARD8, IAN WONG9, JEAN-MICHEL DÉSERT10, P.R. MCCULLOUGH11, ELIZA M.-R.
KEMPTON12,13,, JONATHAN FORTNEY14, RONALD GILLILAND15, DRAKE DEMING12, JOSHUA
KAMMER16
[1] Department of Physics and Institute for Research on Exoplanets, Université de Montréal, Montréal, QC, Canada
[2] Division of Geological and Planetary Sciences, California Institute of Technology, Pasadena, CA 91125, USA
[3] Lunar & Planetary Laboratory, University of Arizona, 1629 E. University Boulevard., Tucson, AZ, USA
[4] Department of Physics and Kavli Institute of Astronomy, Massachusetts Institute of Technology, 77 Massachusetts
Ave, Cambridge, MA, 02139, USA
[5] Space Science Institute, 4750 Walnut Street, Suite 205, Boulder, CO 80301, USA
[6] Department of Astronomy, University of Texas, Austin, TX 78712, USA
[7] Department of Astronomy, Harvard University, 60 Garden Street, Cambridge, MA 02138, USA
[8] Department of Astronomy, California Institute of Technology, Pasadena, CA 91125, USA
[9] Department of Earth, Atmospheric, and Planetary Sciences, Massachusetts Institute of Technology, 77 Massachusetts
Ave, Cambridge, MA, 02139, USA
[10] Anton Pannekoek Institute for Astronomy, University of Amsterdam, 1090 GE Amsterdam, The Netherlands
[11] Department of Physics and Astronomy, Johns Hopkins University, Baltimore, MD 21218, USA
[12] Department of Astronomy, University of Maryland, College Park, MD 20742, USA
[13] Department of Physics, Grinnell College, 1116 8th Avenue, Grinnell, IA 50112, USA
[14] Department of Astronomy, University of California, Santa Cruz, CA 95064, USA
[15] Space Telescope Science Institute, 3700 San Martin Dr., Baltimore, MD 21218, USA
[16] Southwest Research Institute, San Antonio TX, USA
[17] IPAC-NASA Exoplanet Science Institute Pasadena, CA 91125, USA
[18] NASA Hubble Fellow
Instrument Filter/Grism Transit/Eclipse Wavelength [µm] UT Start Date
2015 Feb 07
HST/STIS G750L Transits 0.55 – 1.0 2015 May 12
2016 Apr 03
2015 Jan 28
HST/WFC3 G141 Transits 1.1 – 1.7 2015 Mar 13
2015 Oct 22
2012 Dec 22
Spitzer/IRAC Channel 1 Transits 3.0 – 4.0 2017 Jan 25
2017 Feb 20
2012 Jun 11
Spitzer/IRAC Channel 2 Transits 4.0 – 5.0 2012 Jun 15
2013 Jan 01
2014 Jan 15
2014 Jan 28
2014 Jun 14
2014 Jun 24
2015 Jan 30
Spitzer/IRAC Channel 1 Eclipse 3.0 – 4.0
2015 Feb 06
2015 Feb 09
2015 Feb 12
2015 Jun 19
2015 Jul 19
2014 Jan 21
2014 Feb 4
2014 Jun 21
2014 Jul 10
2015 Jan 10
Spitzer/IRAC Channel 2 Eclipse 4.0 – 5.0
2015 Jan 13
2015 Jan 17
2015 Jan 20
2015 Jan 23
2015 Jan 27
Supplementary Table 1: Summary of presented transit and eclipse observation of GJ 3470b.
Supplementary Figure 1: White light curve fit (left) and a typical spectral light curve fit (right) from the
joint analysis of the three WFC3 transit observations of GJ 3470b. The top panel shows the best fitting
model light curves (black curve), overlaid with the systematics-corrected data (circles). Residuals from the
light curve fits are shown in the middle panels. All corrected WFC3 light curve fits are free of obvious
systematics. The bottom panels shows a histogram of the residuals normalized by the fitted photometric
scatter parameter for each respective transit. The residuals follow the expected Gaussian distribution for
photon noise limited observations.
Supplementary Figure 2: White light curve fit (left) and a typical spectral light curve fit (right) from the
joint analysis of the three STIS transit observations of GJ 3470b. The top panel shows the best fitting model
light curves (black curve), overlaid with the systematics-corrected data (circles). Residuals from the light
curve fits are shown in the middle panels. The bottom panels shows a histogram of the residuals normalized
by the fitted photometric scatter parameter for each respective transit.
Supplementary Figure 3: Spitzer light curve fits of three 3.6µm transit (left) and three 4.5µm transits (right).
The top panel shows the best fitting model light curves (black curve), overlaid with the systematics-
corrected data (colored circles). Residuals from the light curve fits are shown in the middle panels. All
corrected Spitzer light curve fits are free of obvious systematics. The bottom panels shows a histogram of
the residuals normalized by the fitted photometric scatter parameter for each respective transit. The
residuals follow the expected Gaussian distribution for photon noise limited observations.
(a) (b)
(e)

(c) (d)

(e)

Supplementary Figure 4: (Left and center) Repeatability of transit depth measurements. Panels (a)-
(d) show the transit depths from the individual transit fits (blue) and joint fit (black) for HST/WFC3
(a), HST/STIS (b), Spitzer/IRAC 3.6µm (c) and Spitzer/IRAC 4.5µm (d). The individual transit
depth measurement are consistent over time within their statistical uncertainties and with the joint
fit. (Right) Modeled spectra showing the potential effects of star spots on the apparent transmission
spectrum. Colored curves indicate the effect for a fraction f of 0.06 (purple), 0.08 (blue), 0.1
(black), 0.12 (green) and 0.13 (red) as discussed in the Methods Section. The second value
indicates the increase in the apparent transit depth within the WFC3 bandpass in parts-per-million
(p.p.m).
Visit l UT Start Date ttrim nbin a Varying Noise Pixel Average Bkd
# (µm) (hr)a aperture? Scaling
rphot a (%)b
1 3.6 UT 2012 Dec 22 0.5 128 no 2.7 0.91
2 3.6 UT 2017 Jan 25 0.75 64 no 2.8 1.01
3 3.6 UT 2017 Feb 20 0.75 128 no 2.5 1.04
1 4.5 UT 2012 Jun 11 0.75 128 no 2.6 0.29
2 4.5 UT 2012 Jun 15 0.75 64 no 2.7 0.32
3 4.5 UT 2013 Jan 01 0.5 64 no 2.7 0.35
1 3.6 UT 2014 Jan 15 1.5 64 yes 1.1x scaling 2.6 0.92
2 3.6 UT 2014 Jan 28 0.5 192 no 2.3 0.82
3 3.6 UT 2014 Jun 14 0.5 128 no 2.2 0.65
4 3.6 UT 2014 Jun 24 1.5 128 yes 0.9x scaling 2.5 0.67
5 3.6 UT 2015 Jan 30 1.5 64 no 2.2 0.69
6 3.6 UT 2015 Feb 06 1.5 128 yes 1.2x scaling 3.0 1.26
7 3.6 UT 2015 Feb 09 0.5 192 no 2.7 1.16
8 3.6 UT 2015 Feb 12 1.5 192 no 2.3 0.91
9 3.6 UT 2015 Jun 19 0.5 128 no 2.2 0.71
10 3.6 UT 2015 Jul 19 0.5 192 no 2.3 0.80
1 4.5 UT 2014 Jan 21 1.5 64 no 2.5 0.30
2 4.5 UT 2014 Feb 04 1.0 128 no 2.3 0.22
3 4.5 UT 2014 Jun 21 1.0 64 no 2.3 0.10
4 4.5 UT 2014 Jul 10 1.0 64 no 2.3 0.29
5 4.5 UT 2015 Jan 10 1.5 64 no 2.1 0.37
6 4.5 UT 2015 Jan 13 1.0 128 no 2.3 0.34
7 4.5 UT 2015 Jan 17 1.5 64 no 2.1 0.30
8 4.5 UT 2015 Jan 20 1.0 128 no 2.4 0.35
9 4.5 UT 2015 Jan 23 0.5 128 no 2.3 0.28
10 4.5 UT 2015 Jan 27 1.5 64 no 2.7 0.34
Supplementary Table 2: Summary of Spitzer transit observations (top) and eclipse observation (bottom)
a
ttrim is the amount of time trimmed from the start of each time series, nbin is the bin size used in the
photometric fits, and rphot is the radius of the photometric aperture in pixels.
b
Relative sky background contribution to the total flux in the selected aperture
Supplementary Figure 5: Spitzer/IRAC secondary eclipse observations of GJ 3470b at 3.6 µm (top) and
4.5 µm (bottom). The left panels show the estimates of the eclipse depths for each of the ten individual
eclipse observations (blue) and the global fit (black). Typical ±1σ uncertainties for individual eclipse fits
are shown by gray dashed horizontal lines around the global fit value. The ten individual measurements are
randomly distributed around the global fit. Consistent with the uncertainties, 6 of 10 and 7 of 10 data points
are within the 68% confidence interval for the observations at 3.6 and 4.5 µm, respectively. The right panels
shows the best fitting model light curves (red curve) from the global fit, overlaid with the systematics-
corrected Spitzer data from all ten eclipse observations (black). The eclipse is slightly offset relative to the
estimated eclipse time for a perfectly circular orbit. This is consistent with fits to the radial velocity data of
GJ 3470b, which independently confirm a small but non-zero eccentricity (Kosiarek et al. 2018).
Instrument/Grism Wavelength Depth +1σ -1σ
[µm] [ppm] [ppm] [ppm]
Transit:
HST STIS G750L 0.528 – 0.577 5912 178 192
HST STIS G750L 0.577 – 0.626 6135 104 104
HST STIS G750L 0.626 – 0.674 5999 89 94
HST STIS G750L 0.674 – 0.723 5945 91 91
HST STIS G750L 0.723 – 0.772 5937 98 107
HST STIS G750L 0.772 – 0.821 6215 85 86
HST STIS G750L 0.821 – 0.870 6093 108 110
HST STIS G750L 0.870 – 0.919 6098 113 119
HST WFC3 G141 1.120 – 1.150 6101 34 35
HST WFC3 G141 1.150 – 1.180 6050 37 34
HST WFC3 G141 1.180 – 1.210 6077 33 34
HST WFC3 G141 1.210 – 1.240 6035 31 32
HST WFC3 G141 1.240 – 1.270 6126 34 31
HST WFC3 G141 1.270 – 1.300 6082 31 29
HST WFC3 G141 1.300 – 1.330 6055 29 29
HST WFC3 G141 1.330 – 1.360 6196 26 28
HST WFC3 G141 1.360 – 1.390 6178 31 30
HST WFC3 G141 1.390 – 1.420 6111 32 32
HST WFC3 G141 1.420 – 1.450 6127 29 31
HST WFC3 G141 1.450 – 1.480 6136 31 32
HST WFC3 G141 1.480 – 1.510 6124 32 31
HST WFC3 G141 1.510 – 1.540 6079 26 29
HST WFC3 G141 1.540 – 1.570 6121 31 32
HST WFC3 G141 1.570 – 1.600 6057 34 30
HST WFC3 G141 1.600 – 1.630 6059 33 35
HST WFC3 G141 1.630 – 1.660 6061 32 32
Spitzer/IRAC 3.6µm 3.15 – 3.90 5763 64 65
Spitzer/IRAC 4.5µm 4.00 – 5.00 5941 60 66

Eclipse:
Spitzer/IRAC 3.6µm 3.15 – 3.90 115 27 26
Spitzer/IRAC 4.5µm 4.00 – 5.00 3 22 22
Supplementary Table 3: Transmission spectrum and eclipse depths
Supplementary Figure 6: Spectral fits and temperature profile constraints from the joint retrieval analysis
of GJ 3470b’s transit and eclipse data. The top and bottom left panels show the range of models fitting the
observations (black) by depicting a random sample of 300 atmospheric models from the posterior
distribution (thin blue curves). The best fitting model is shown as thick red curve. Consistent with the
observed transit depth uncertainties, the models are tightly constrained within the precise HST/WFC3
observations. At shorter and longer wavelengths, the HST/STIS and Spitzer observations allow for a
slightly larger range of transit depths, mostly by slight changes in the cloud parameters as well as the CO
and CO2 abundances. The bottom right panel depicts the posterior constraints on the vertical temperature
profile. The dark and light-blue shaded regions are the 1σ and 2σ spread in the temperature profiles, with
the solid blue curve being the median temperature profile of all models in the posterior distribution. The
equilibrium temperature for an planetary albedo of 0.1 is shown for comparison (dashed line).
Supplementary Figure 7: Molecular abundance and cloud property constraints from the joint retrieval
analysis of the transit and eclipse data. The top panels in each column show the 1D marginalized posterior
distributions of the molecular abundances and cloud properties, with dashed vertical lines in the
histograms indicating the marginalized 16th, 50th, and 84th percentiles. The subjacent 2D panels show
the correlations among the gases and cloud properties, with the black, dark-gray, and light-gray regions
corresponding to the 1σ (39.3%), 2σ (86.5%), and 3σ (98.9%) credible intervals. The vertical and
horizontal red lines in each panel are the solar composition molecular abundances at 700 K and 0.1 bars, a
representative photospheric temperature and pressure. The water mixing ratio is constrained to ±1 order
of magnitude around 1 times solar. CH4 and NH3 are depleted. Note also the “elbow”-shaped correlation
between CO and CO2. This degeneracy arises because CO and CO2 both absorb within the 4.5 µm Spitzer
bandpass observed in transit and eclipse. Note that the retrieval included an additional 7 parameters for
the vertical temperature structure and common-mode transit depth uncertainties which are not displayed
here for clarity.
Supplementary Figure 8: Mixing-ratio profiles for several species of interest (as labeled) in our
kinetics/transport models for GJ 3470b, for solar atmospheric metallicity and a C/O ratio of 0.54. The
gray horizontal zone indicates pressure range to which our observations are most sensitive. The thermal
emission observations extend to slightly deeper levels as well. Water and methane abundance follow
mostly the equilibrium abundances, with photodissociation becoming relevant above approximately 10-5
bar, as previously also shown in a theoretical modeling investigation of the sub-Neptune GJ 1214b15 and
GJ 436b16,21. Ammonia is expected to be abundant the photosphere due to quenching resulting from
vertical transport and the slow rate of reaction for the conversion to N2 and H2.
Supplementary Figure 9: Analysis of residuals from fitting the three Spitzer/IRAC 3.6um transits (top,
center, bottom). Left panels: Photometric scatter vs. the width of the binning interval for Spitzer data. The
root-mean-square error of the systematics-corrected Spitzer data (black) follows closely the theoretical
square-root scaling for uncorrelated white noise (red dashed line), even when binned all the way to 30
minute intervals. Right panels: Histogram of the residuals (grey bars) compared to a theoretical Gaussian
distribution with the width of the scatter parameter fitted as a nuisance parameter in the Bayesian analysis
(black curve). The residuals are consistent with the Gaussian distribution and the scatter parameter.
Supplementary Figure 10: Analysis of residuals from fitting the three Spitzer/IRAC 4.5um transits (top,
center, bottom). Left panels: Photometric scatter vs. the width of the binning interval for Spitzer data. The
root-mean-square error of the systematics-corrected Spitzer data (black) follows closely the theoretical
square-root scaling for uncorrelated white noise (red dashed line), even when binned all the way to 30
minute intervals. Right panels: Histogram of the residuals (grey bars) compared to a theoretical Gaussian
distribution with the width of the scatter parameter fitted as a nuisance parameter in the Bayesian analysis
(black curve). The residuals are consistent with the Gaussian distribution and the scatter parameter.
Supplementary Figure 11: Histograms of residuals from HST/WFC3 spectral light curve fits. This figure
shows the histograms for each wavelength bin (rows) and each transit (columns) individually. Residuals
are normalized by the fitted scatter parameter for the respectively wavelength bin and transit. The histogram
of normalized residuals (gray bars) is compared to the normal distribution (black curve). Each histogram is
made of only 51 data points leading to relatively poor sampling of each frequency distribution; however,
no statistically significant deviation from the expected frequency distribution is observed. The agreement
with the expected Gaussian distribution can be seen even better in the combined plot of all residuals shown
in Supplementary Figure 12. WFC3 residuals are highly consistent with the Gaussian distribution and the
fitted scatter parameter for each light curve.
Supplementary Figure 12. Histograms of all residuals from WFC3 spectral light curve fits. This plot
combines the residuals from all panels in Supplementary Figure 11 in order to increase the number of
samples in the histogram. WFC3 residuals are highly consistent with the Gaussian distribution and the fitted
scatter parameter for each light curve.
Supplementary Figure 13: Photometric scatter vs. the width of the time binning interval for all WFC3
spectroscopic light curves combined. The root-mean-square error of the systematics-corrected WFC3 data
(black) follows closely the theoretical square-root scaling for uncorrelated white noise (red dashed line),
even when binned all the way to 12 minute intervals. At this point only four data points are left per orbit.
We conclude that time correlated noise is negligible.
Supplementary Figure 14: Pairs plot showing the posterior distribution of the MCMC fitting parameters for
the WFC3 spectral light curve fit. The panels on the diagonal show the marginalized posterior distribution
for each fitting parameter. The 68% credible interval is marked by vertical dashed lines and quantified
above the panel. The off-diagonal panels show the two-dimensional marginalized distribution for pairs of
parameters, with the gray shading corresponding to the probability density and black contours indicating
the 68% and 95% credible regions. Our instrument modeling results in no significant correlation between
astrophysical transit depth (first column) and instrumental detrending parameters.
Supplementary Figure 15: Histograms of residuals from HST/STIS spectral light curve fits. This figure
shows the histograms for each wavelength bin (rows) and each transit (columns) individually. Residuals
are normalized by the maximum likelihood value of the scatter parameter for the respectively wavelength
bin and transit. The histograms of the normalized residuals (gray bars) are compared to the normal
distribution (black curve). Each histogram is made up of 29 data points leading to relatively poor sampling
of each frequency distribution. All residuals combined are shown in Supplementary Figure 16. STIS
residuals are consistent with a Gaussian distribution and the fitted scatter parameter is a conservative
estimate of the scatter.
Supplementary Figure 16: Histograms of all residuals from STIS spectral light curve fits. This plot
combines the residuals from all panels in Supplementary Figure 15 in order to improve the number of
samples in the histogram. A distribution marginally narrower than the median of the scatter parameters is
found. We conclude that our Bayesian analysis of the STIS light curves conservatively estimated the error
bar as a result of the many detrending parameters needed to fit STIS light curves. This results in a
conservative estimate of the transit depth uncertainties. Note that a standard maximum likelihood method
would have found a smaller scatter because it would have estimated the scatter only based on the best fitting
(potentially overfitting) model.
Supplementary Figure 16: Photometric scatter vs. the width of the binning interval for the ten Spitzer/IRAC
3.6um eclipses. The root-mean-square error of the systematics-corrected Spitzer data (black) follows
closely the theoretical square-root scaling for uncorrelated white noise (red dashed line), even when
combining up to 100 to 1000 data points to one bin.
Supplementary Figure 17: Photometric scatter vs. the width of the binning interval for the ten Spitzer/IRAC
3.6µm eclipses. The root-mean-square error of the systematics-corrected Spitzer data (black) follows
closely the theoretical square-root scaling for uncorrelated white noise (red dashed line), even when
combining up to 100 to 1000 data points to one bin.

You might also like