This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.

CHARACTERIZING TRANSITING EXOPLANET ATMOSPHERES WITH JWST

, , , , , and

Published 2016 January 19 © 2016. The American Astronomical Society. All rights reserved.
, , Citation Thomas P. Greene et al 2016 ApJ 817 17 DOI 10.3847/0004-637X/817/1/17

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/817/1/17

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.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

There are now well over a thousand confirmed exoplanets, ranging from hot to cold and large to small worlds. Characterizing the atmospheres of a diversity of planets is critical to understanding their bulk compositions, formation (and any migration), energy balance, and atmospheric processes (e.g., see Burrows & Orton 2010, p. 419; Seager & Deming 2010; Burrows 2014; Madhusudhan et al. 2014, p. 739; Crossfield 2015; Heng & Showman 2015).

Much of the atmospheric characterization work has come from the acquisition and interpretation of transmission and emission spectroscopy of transiting planets by the Hubble Space Telescope (HST) and Spitzer space telescope. The transmission or emission spectra of hot Jupiters (e.g., HD189733b, WASP-43b), warm Neptunes (e.g., GJ 436b, HAT-P-11b), and warm sub-Neptunes (e.g., GJ 1214b) are all being studied extensively with HST, Spitzer, and other facilities. The near-IR spectra of cool or warm super-Earths like K2-3b/c/d will soon be observed with HST, and they will likely be prime candidates for James Webb Space Telescope (JWST) spectroscopy. Observations to date have produced a variety of important discoveries such as the detection of H2O absorption, now clearly seen in a variety of planets using a variety of instruments. HST NICMOS and HST WFC3 G141 find strong water absorption in about a dozen hot Jupiters (e.g., Swain et al. 2009; Crouzet et al. 2012; Deming et al. 2013; Kreidberg et al. 2014b, 2015) and a warm Neptune-sized planet (Fraine et al. 2014). Atmospheric retrieval techniques have been applied to the data in order to determine the abundances (or upper limits) of molecules such as CO, CO2, and CH4 in addition to H2O in ∼10 exoplanet atmospheres (e.g., Madhusudhan & Seager 2009; Madhusudhan et al. 2011; Benneke & Seager 2012, 2013; Lee et al. 2012; Line et al. 2012, 2013b, 2014b; Barstow et al. 2013a, 2013b; Benneke 2015; Waldmann et al. 2015). Molecular abundance determinations have been used to constrain atmospheric C/O ratios (e.g., Madhusudhan et al. 2011; Line et al. 2014b; Benneke 2015), which can potentially help diagnose where a planet formed relative to the H2O and CO ice lines in its protoplanetary disk (Öberg et al. 2011). Atmospheric metallicity determinations relative to the host star have also been used to infer formation possibly via core accretion (e.g., Kreidberg et al. 2014b). Temperature inversions have been suggested to explain the emission spectra of a number of hot Jupiters (e.g., see Fortney et al. 2011; Knutson et al. 2010, and references therein), indicating the presence of visible or UV stratospheric absorbers. However, recent work by Line et al. (2014b) indicates no strong statistical evidence for inversions in a sample of nine observed planets, but the very hot Jupiters HAT-P-7b (Christiansen et al. 2010) and WASP-33b (Haynes et al. 2015) do appear to have temperature inversions at the present time (see also Crossfield 2015). Determining the frequency of inversions over a variety of bulk planetary properties is important for understanding chemical processes (e.g., impact of C/O on high-altitude absorbers) and the overall energy balance in these planets' atmospheres.

Despite these advances, there are still considerable uncertainties in the compositions, temperatures, and origins of exoplanet atmospheres. Numerous early HST and Spitzer detections of molecular features and temperature inversions have been called into question or disproven with subsequent higher precision observations, more sophisticated data analysis, and powerful modern retrieval techniques for molecular abundances and temperature–pressure (hereafter TP) profiles (e.g., Gibson et al. 2011; Diamond-Lowe et al. 2014; Line et al. 2014b; Benneke 2015; Schwarz et al. 2015). Clearly, we need to better determine the compositions of exoplanet atmospheres, what planets have stratospheric temperature inversions under what conditions, how closely planet elemental abundances match their host stars, and where planets formed in their disks. Addressing or resolving these specific questions would significantly advance our understanding of exoplanet atmospheres:

  • Do any highly insolated planets have hot stratospheres, and what absorbers are causing these temperature inversions if they exist?
  • What are the natures of super-Earth (1.5–2 R) planet atmospheres; are they mostly H and He or are they dominated by species of high mean molecular weight (HMMW) like Earth and Venus?
  • How do clouds inhibit our ability to infer molecular abundances?
  • What are the C-to-O ratios in planetary atmospheres, how do these compare to their host stars, and what does this imply about where planets formed in protoplanetary disks?
  • How do the metal abundances of planets compare to their host stars, and does this vary by planet mass as it does for giants in our solar system?
  • How far from chemical equilibrium can exoplanet atmospheres be driven, and what causes this?

High-quality JWST observations may characterize transiting exoplanet atmospheres well enough to address these questions significantly in the near future (launch is currently scheduled for 2018 October). JWST's large aperture (6.5 m), numerous spectroscopic modes over λ = 0.6–28 μm, good thermal stability, and applications of lessons learned from other observatories will ensure that it collects the highest quality exoplanet transmission and emission spectra. Numerous studies are providing assessments of how well JWST is expected to characterize exoplanet atmospheres. Beichman et al. (2014) present many details of JWST's instruments and recommend appropriate modes for observing transiting exoplanets. Cowan et al. (2015) report that JWST is expected to characterize dozens of giant planets over its mission lifetime, but observations of cool, small-planet ("temperate terrestrial") atmospheres may require ∼100 days each. Batalha et al. (2015) find that JWST NIRSpec should be able to measure the λ = 1–5 μm transmission spectra of nearby (3–50 pc), cool (400–1000 K), low-mass (1–10 M) planets orbiting mid-M dwarfs with moderate signal-to-noise ratios (S/Ns) after summing 25 transits. Barstow et al. (2015) performed JWST data simulations and atmospheric retrievals to investigate the impacts of starspots and other systematic errors that shift different spectral wavelength ranges that are not obtained simultaneously.

These studies have greatly improved our understanding of the promises and potential limits of JWST data. However, we still do not understand how well JWST will be able to quantitatively characterize different types of planets in the above ways. We also need to understand better what observing modes will be most useful for addressing specific questions. This is an important assessment because most transiting planets with bright host stars (J ≲ 11 mag) will need to be observed four separate times to collect their entire λ = 0.7–11+ μm spectra (e.g., see Beichman et al. 2014), and JWST time will be extremely precious. We perform and discuss such quantitative assessments in this contribution.

We investigate how well constraints on temperature and molecular volume mixing ratios from JWST observations can address these big-picture questions by analyzing simulated observations of a diverse range of planet types. We start by presenting a diverse set of planetary systems that span what we think are the typical planet types in Section 2. We describe the atmospheric models and the retrieval technique in Section 3. Next we describe the instrument signal and noise models we use to create simulated JWST spectra in Section 4. The retrieval results including the molecular and temperature constraints are presented in Section 5. We apply these results to assess planet carbon-to-oxygen ratios, metallicities, and probes of disequilibribum chemistry in Section 6 and the big-picture questions these can address. Finally, we summarize our conclusions in Section 7.

2. SIMULATED PLANETS

We assemble a set of four fiducial planetary systems to assess how well JWST will be able to characterize exoplanet atmospheres in the early years of its mission. These planets range from hot (Teq = 1500 K) to cool (Teq = 500 K) and large (1.36 RJ) to small (0.19 RJ). This combination of sizes and temperatures maps well onto the now established planet archetypes of hot Jupiters, warm Neptunes, warm sub-Neptunes, and cool super-Earths. We select the physical parameters of a well known system from each of these categories to assemble our set of four systems to model. We model their atmospheres as either clear with solar elemental abundances, cloudy with solar elemental abundances, and for the smaller planets clear with enhanced elemental abundances so as to result in HMMW atmospheres.

Molecular abundances that are constant with altitude were generated assuming broad consistency with thermochemical equilibrium given the effective temperature of the planet and the elemental abundances (computed with the Chemical Equilibrium with Applications code; Gordon & McBride 1994; Line et al. 2010; Moses et al. 2011; Visscher & Moses 2011)9 , and the HMMW atmospheres were either 1000 × solar metallicity or pure H2O. The 1000× solar compositions correspond to mean molecular weight μ = 15 for the warm Neptune and μ = 16.8 for the warm sub-Neptune atmospheres (see Fortney et al. 2013). It is not clear whether 1000 × solar metallicity is possible since accreted planetesimals include hydrogen if they are icy (Fortney et al. 2013), preventing this level of atmospheric metallicity. Nevertheless, we believe that this is a useful bounding case. The 100% H2O cool super-Earth atmosphere has a molecular weight of μ = 18 and corresponds to an [Fe/H] = 2.77 (588 × solar metallicity).10 The sets of planet types, system parameters, and atmospheric molecular mixing ratios are given in Tables 13. Actual small, cool planets may not be in chemical equilibrium, and we discuss how well we can detect disequilibrium chemistry in Section 6.3.

Table 1.  Planetary Systems Modeled

Planet Type System Parameters Composition Clouds Geometry
Hot Jupiter HD 209458b 1×solar Clear Trans., Emis.
      1 mbar Trans
Warm Neptune GJ 436b 1×solar Clear Trans., Emis.
      1 mbar Trans
    1000× solar Clear Trans., Emis.
Warm sub-Neptune GJ 1214b 1×solar Clear Trans., Emis.
      1 mbar Trans
    1000× solar Clear Trans., Emis.
Cool super-Earth K2-3b 1×solar Clear Trans., Emis.
      1 mbar Trans
    100% H2O Clear Trans., Emis.

Download table as:  ASCIITypeset image

Table 2.  Fiducial Planetary System Parameters

Planet Type System Parameters T* (K) R* (R) K (mag) Teqa (K) Mp(M) Rp(R) Hb (km) T14 (s)
Hot Jupiter HD 209458b 6065 1.155 6.3 1500 220 15 560 11,000
Warm Neptune GJ 436b 3350 0.464 6.1 700 23 4.2 190 2740
Warm sub-Neptune GJ 1214b 3030 0.211 8.8 600 6.5 2.7 230 3160
Cool super-Earth K2-3b 3900 0.561 8.6 500 5.3c 2.1 150c 9190

Notes. Tabulated system values were taken from the exoplanets.org compilation (Han et al. 2014) and Crossfield et al. (2015).

aEquilibrium temperature Teq was computed from the listed system values assuming albedo = 0 and energy redistribution over 4π sr. bThe scale height of the planetary atmosphere H = kTeq/(μmHg) for the clear solar atmosphere of each planet (μ = 2.3) is provided as a convenience for scaling to other systems. cThe mass of this planet has been recently measured to be 8.4 ± 2.1 M (Almenara et al. 2015), somewhat higher than the tabulated value we used in our investigation. This increased mass would decrease the scale height, decrease the S/N of transmission spectral features, and worsen the derived abundance precisions by roughly 40%.

Download table as:  ASCIITypeset image

Table 3.  Molecular Volume Mixing Ratios of Planetary Atmospheres

Planet Type Composition H2O CH4 CO CO2 NH3 N2
Hot Jupiter solar 4.27 × 10–4 1.00 × 10–9 4.27 × 10–4 1.26 × 10–7 3.16 × 10–10 5.75 × 10–5
Warm Neptune solar 7.24 × 10–4 4.27 × 10–4 1.00 × 10–9 3.16 × 10–11 3.16 × 10–5 2.51 × 10–5
Warm Neptune 1000× solar 2.51 × 10–1 1.00 × 10–1 1.00 × 10–2 1.45 × 10–1 5.01 × 10–5 4.47 × 10–2
Warm sub-Neptune solar 7.24 × 10–4 4.26 × 10–4 1.00 × 10–9 3.16 × 10–11 3.16 × 10–5 2.51 × 10–5
Warm sub-Neptune 1000× solar 3.98 × 10–1 1.26 × 10−1 1.00 × 10–3 1.26 × 10–1 5.01 × 10–5 5.01 × 10–2
Cool super-Earth solar 7.24 × 10–4 4.27 × 10–4 1.00 × 10–11 1.00 × 10–11 1.00 × 10–4 1.58 × 10–5
Cool super-Earth 100% H2O 1.00 0 0 0 0 0

Note. All mixing ratios are assumed constant with altitude and are thermochemical equilibrium values for the planet temperatures and compositions given in Tables 1 and 2 except that the hot-Jupiter terminator temperature of 1200 K (Moses et al. 2011) was used instead of its Teq value in Table 2.

Download table as:  ASCIITypeset image

3. MODELING AND RETRIEVAL APPROACH

The emission spectra are computed using the forward model described in Line et al. (2013b) 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. (2013a), Swain et al. (2014), and Kreidberg et al. (2014b). We then use a derivative of the CHIMERA11 retrieval suite (Line et al. 2013b, 2013a) to determine the degree of constraints on temperatures and abundances from simulated observations of the transmission and emission spectra (see Section 4). The molecular opacities described in Section 2.3 of Line et al. (2015) were used for computing all forward and retrieval model spectra.

While atmospheres are undoubtably complicated, we choose a relatively simple 1D parameterization. Since we are using the same forward model to generate the synthetic spectra as the retrieval, this should not be a problem for assessing the impact of JWST data quality. This will allow us to strictly explore the role that the JWST instrumental noise properties and spectral coverage have in setting the atmospheric constraints. We perform a preliminary exploration of retrieval assumptions and priors and their inherent biases on the retrieved results later in this work (see Section 5.1), and a more thorough analysis will be the subject of a forthcoming paper.

Because of the different viewing geometries, the atmospheres used in transmission are parameterized differently than the atmospheres used to generate the emission spectra. However, both use the same set of molecular abundances that are uniform with altitude. These include H2O, CH4, CO, CO2, NH3, and N2 (except in emission as it has no consequence on the emission spectra). Any remaining gas is assumed to be a mixture of solar composition H2/He. N2 is the dominant nitrogen-bearing species at high temperatures, but it has no spectroscopic features unless in high concentrations (Schwieterman et al. 2015). We include it here as a trace gas simply to contribute to the mean molecular weight of atmospheres in transmission spectra. The fiducial abundances are again chosen to be broadly consistent with thermochemical equilibrium at the given scale height temperature. For all but the hot-Jupiter scenario we deplete two oxygen atoms per magnesium atom to emulate oxygen loss due to enstatite (Mg2SiO4) condensation. The impact of various absorption features for each of these species is shown in the transmission and emission model spectra for a hot Jupiter and warm Neptune of clear solar composition in Figure 1.

Figure 1.

Figure 1. Left: magnitude of absorption for specific gases for the transmission (upper left) and emission (lower left) models for a hot Jupiter of clear solar composition. Right: wavelengths and amplitudes of gaseous absorptions for the transmission (upper right) and emission (lower right) models for a warm Neptune of clear solar composition. Each panel shows the impact of the dominant molecules in each model.

Standard image High-resolution image

The transmission spectra are generated with 11 free parameters, six of which are the volume mixing ratios of the molecular gasses noted above. The remaining five parameters are as follows: first, an effective scale height temperature (T). This will directly impact the amplitude of the spectral features. For simplicity, we assume isothermal atmospheres for the transmission spectra at the approximate equilibrium temperature (without truly knowing the albedo or redistribution) of the planet in question. Barstow et al. (2013b, 2015) demonstrated that there may be potential biases in real atmospheres if isothermal atmospheres are assumed, but that it will be incredibly difficult to actually retrieve detailed temperature profile information from transmission spectra. Second, we use a scaling to the fiducial 10 bar planet radius (xRp). This parameter manifests itself as a zero-point offset in the spectra as well as a minor impact on the amplitude of the features. Third, we include an opaque gray cloud parameterized with a cloud-top pressure (Pc) that we set to 1 mbar. The atmospheric transmittance at atmospheric levels deeper than the cloud top is set to zero. Several measured flat exoplanet transmission spectra have been reasonably well described with such a parameterization (Line et al. 2013a; Knutson et al. 2014; Kreidberg et al. 2014a). We also model hazes in an ad-hoc fashion using the following approximation (e.g., Lecavelier Des Etangs et al. 2008):

Equation (1)

σ0 is the magnitude of the haze cross-section relative to H2 Rayleigh scattering at 0.4 μm, and β is the wavelength power-law index that describes the slope. While the λ ≥ 1 μm wavelength range explored in this study will not provide constraints on these parameters, we include them because their degeneracy with the absorbers at near-IR wavelengths could potentially influence the retrieved abundances (e.g., Benneke 2015; Kreidberg et al. 2015). Figure 2 shows the spectra for all planets modeled in transmission (see also Table 1).

Figure 2.

Figure 2. Forward model transmission spectra of all planets in Table 1 where Absorption Depth = (Rp,λ/R*)2. Spectra have been binned to resolution R ≤ 100 (hot Jupiter, warm Neptune, warm sub-Neptune; see Section 4.2) or R = 35 (cool super-Earth) to match that shown for the simulated JWST data. Dashed lines show the wavelength range boundaries of the chosen NIRISS, NIRCam, and MIRI instrument modes. The vertical offsets in the cloudy spectra are due to increased opacity from opaque clouds, increasing the effective planetary radii (e.g., see Brown 2001). The smaller absorption depths of the HMMW atmospheres are due to their smaller (than solar composition) scale heights. (The data used to create this figure are available.)

Standard image High-resolution image

The emission spectrum model requires 10 free parameters (five of which are the molecular absorbers). We assume a thermally and chemically homogenous dayside, represented by a single 1D TP profile. Because thermal emission spectra contain temperature "profile" information, we use a five-parameter double-gray analytic formula (Line et al. 2013b, and references therein). We assume cloud-free emission spectra. Although clouds can dominate transmission spectra, they may have a lessened impact on emission spectra due to their reduced effective optical depths (Fortney 2005). The specific impact of clouds on emission spectra depends on where (at what pressure) the thermal emission arises within the atmosphere, where clouds form (i.e., where their material condensation curves cross atmospheric TP profiles), and the absorption and scattering properties of all clouds within the pressure region probed by the emission. This is complex enough to warrant its own investigation, and we do not explore the impact of clouds on emission spectra in this work. HST WFC3 dayside spectra of several hot Jupiters are consistent with cloud-free atmospheres (McCullough et al. 2014; Stevenson et al. 2014; Kataria et al. 2015), so we do know that clouds do not suppress spectral features in emission spectra of all planets. Figure 3 shows the spectra of all planets modeled in emission (see also Table 1).

Figure 3.

Figure 3. Forward model emission spectra of all planets in Table 1. Spectra have been binned to resolution R ≤ 100 (hot Jupiter, warm Neptune, warm sub-Neptune; see Section 4.2) or R = 35 (cool super-Earth) to match that shown for the simulated JWST data. Dashed lines show the wavelength range boundaries of the chosen NIRISS, NIRCam, and MIRI instrument modes. (The data used to create this figure are available.)

Standard image High-resolution image

More complicated atmospheric processes such as vertical mixing, atmospheric dynamics, photochemistry, ion chemistry (both photochemical and thermal), unknown cloud processes, and perhaps many other known or unknown processes (all active areas of research) could make our chosen parameter values less than realistic. Certainly any of these processes could be modeled with a nearly infinite number of unknown parameters. However, we are only investigating the impacts of the spectral (wavelengths and resolutions) and noise properties of JWST instruments on retrieving information from their observations. Therefore these assumptions are perfectly appropriate for this task of assessing the uncertainties of retrievals and any systematic deviations from the adopted forward models. We will use these parameters and their constraints to evaluate the expected JWST performance.

We assume uniform or uniform-in-log priors12 for the gases and other parameters, as done in Line et al. (2014a), Stevenson et al. (2014), and Kreidberg et al. (2014b). The retrievals are initialized at the true parameter vector to minimize burn-in and to speed convergence.

4. SIMULATED JWST SPECTRA

We have chosen to simulate the λ = 1–11 μm transmission and emission spectra of the modeled systems. This region of the spectrum shows the dominant absorption features for the major carbon-, oxygen-, and nitrogen-bearing species. We do not focus on constraining alkali metals and metal oxides/hydrides in this investigation. These species (for hot planets), along with molecular Rayleigh scattering and possible hazes, will dominate the spectrum below 1 μm.

We now describe our simulations of JWST spectra of the planetary system forward models described in Section 2 (Tables 1 and 2) and Section 3. As described in Beichman et al. (2014), all four JWST science instruments have spectroscopic modes that will likely be useful for observing transiting planets. Additional information on expected instrument performance and spectroscopic observations of transiting planets has been published for all instruments: NIRISS (Doyon et al. 2012), NIRSpec (Ferruit et al. 2014; Batalha et al. 2015), NIRCam (Greene et al. 2007), and MIRI (Kendrew et al. 2015). The flexible capabilities of its instruments will allow JWST users to often choose between more than one instrument mode for obtaining a spectrum of a given wavelength range. As noted in Section 1, observations over a large spectral range with low to moderate spectral resolution (R ≡ λ/δλ ≳ 35) and high spectro-photometric precision have provided the best scientific constraints on transiting exoplanet atmospheres to date. Slitless spectra with good spatial sampling and stable detectors have provided the best spectro-photometric precision (e.g., HST WFC3 G141 mode as in Kreidberg et al. 2014a).

We have chosen to simulate λ = 1–11 μm transmission and emission spectra of the modeled systems for the instrument modes listed in Table 4. We selected these modes because they best meet the criteria noted above: large simultaneous wavelength coverage, adequate spectral resolution (R ≳ 100), slitless operation, and bright limits sufficient to observe the selected systems with high throughput. They also have the best spatial sampling available on JWST over their wavelength ranges; good sampling should minimize systematic errors due to variations in intrapixel response in the presence of pointing jitter (e.g., Deming et al. 2009). The NIRSpec instrument can also obtain spectra of bright objects over the same wavelengths as NIRISS SOSS and the NIRCam LW grisms (also in three exposures), and its use in exoplanet transit observations has been well studied (e.g., Beichman et al. 2014; Ferruit et al. 2014; Batalha et al. 2015). However, we chose to simulate NIRISS and NIRCam over 1–5 μm wavelengths because of their slitless operation, finer spatial sampling (64 versus 100 mas pixel−1), and brighter flux limits (important for the best targets).

Table 4.  Selected JWST Instrument Modes

Instrument Mode Optics λ (μm) Native Ra Sampling (pixels)b
NIRISS bright SOSS GR700XD 1–2.5c ∼700 ∼25
NIRCam LW grism F322W2 2.5–3.9 ∼1700 ∼2
NIRCam LW grism F444W 3.9–5.0 ∼1700 ∼2
MIRI slitless LRS prism 5.0–11d ∼100 ∼2

Notes.

aSpectral resolution R ≡ λ/δλ. bSpatial extent of point-source spectrum in pixels. cTotal NIRISS SOSS mode wavelength coverage spans λ = 0.6–2.8 μm, but we adopt the 1.0 μm short-wavelength cutoff that is required for these bright stars and a 2.5 μm long-wavelength cutoff to avoid spectral contamination. dWe adopt a long-wavelength cutoff of 11 μm for the MIRI LRS because its transmission and S/N are degraded at longer wavelengths (see Figures 8 and 9 of Kendrew et al. 2015).

Download table as:  ASCIITypeset image

4.1. Signals and Random Noise Components

Signals and noise of transmission and eclipse spectra were estimated for each instrument mode (Table 4) for each planetary system (Table 1). We combined these synthetic spectra to mimic the observational sequence of observing a single transit or eclipse event with each instrument mode (a total of four events for each transit or eclipse), producing a complete simulated λ = 1–11 μm spectrum and error bars estimating the uncertainty in each spectral channel.

We computed signals for the separate stages of a transit or eclipse event: star only or star + planet. Model stellar flux spectra were constructed by interpolating Nextgen (Hauschildt et al. 1999) models with surface gravities and effective temperatures spanning the observed values for each system. The in-transit star + planet flux was defined as $F{(\ast +p)}_{\mathrm{Tr},\lambda }={F}_{* ,\lambda }\times (1-{({R}_{p,\lambda }/{R}_{* })}^{2})$. The star + planet emission flux is simply the sum of the model star flux and the model planet flux (see Section 3) at each wavelength, $F{(\ast +p)}_{\mathrm{Em},\lambda }={F}_{* ,\lambda }+{F}_{p,\lambda }$. Estimated detected signals of each of these astrophysical events were computed for each system using the chosen system parameters (Table 4), with

Equation (2)

where Sλ is the signal per spectral element in electrons, Fλ is the spectral flux density of the star or star + planet, t is the total exposure time during the event, λ is the wavelength of the spectral element, R is the resolution of the spectral bin, Atel is the area of the JWST aperture (25 m2), and τ is the total system transmission (photon conversion efficiency) at that wavelength. Total exposure time t was calculated as the sum of the photon collecting time of all integrations that are executed during 0.9× the event duration (T14 in Table 2). Individual integration times were computed from the host star brightnesses (Table 2) and the correlated double-sample minimum integration times and bright limits of each instrument mode (see Beichman et al. 2014), including the resulting readout efficiency.

We assumed that the star would also be observed for a time t before and/or after the transit or eclipse event, yielding total integration time 2t for the visit. The system transmission τ was computed to be the product of that of the three-element JWST telescope (set to 0.9) and that of the selected instrument. The transmission τ of NIRISS single-object slitless spectroscopy (SOSS) was estimated to be 0.35 at λ = 1.25 μm, falling off gradually at longer and shorter wavelengths in an approximation of the first-order blaze function of its new GR700XD grism. Similarly, the NIRCam was estimated to have τ = 0.30 at the 3.7 μm blaze peak of its grisms, falling off at longer and shorter wavelengths in an approximation of the first-order grism blaze function (Greene et al. 2007). The slitless photon conversion efficiency curve (τ ≃ 0.3) in Figure 9 of Kendrew et al. (2015) was adopted for the MIRI LRS τ values.

The JWST observatory is expected to have minimal natural (zodiacal) and telescope background emission at near-IR wavelengths, but the telescope background will increase quickly at wavelengths λ ≳ 10 μm. Background emission can become an important noise term when observing faint stars with the MIRI LRS because its slitless mode detects the full background of its λ = 5–12 μm bandpass. We computed the background signal for each observation with the equation

Equation (3)

where Bkg is the background signal detected in each spectral bin, B is the background of the instrument mode in electrons arcsec−2 s−1, t is the total exposure time during the event, Apix is the area subtended by each pixel in arcsec2, npix is the number of spatial × 2 spectral pixels summed in each Rnative native two-pixel resolution element of the selected observing mode, and R is the final binned spectral resolution of the observation. The B background values were computed from the average zodiacal and telescope backgrounds of each instrument mode provided by the STScI JWST prototype exposure time calculator.13

Noise values were computed at each wavelength as the sums of photoelectron Poisson noise, background Poisson noise, and total (read and dark) detector noise Nd,tot:

Equation (4)

where

Equation (5)

Nd is the total (read and dark current) detector noise of a single integration, and nints is the number of integrations during exposure time t. We set Nd = 18 electrons for the HgCdTe detectors (NIRISS and NIRCam modes) and Nd = 28 electrons for MIRI; these correlated double-sample noise values are expected maxima for nearly all observations. The numbers of integrations nints were set using the bright limits, efficiencies, and frame times of each instrument mode given in Beichman et al. (2014) such that the total real time fit within 0.9 times the transit duration (T14 in Table 2), summing to total exposure time t for the event (plus equal time on the host star alone).

4.2. Final Simulated Spectra

The signals and backgrounds of the astrophysical events were combined as follows to estimate the final transmission (Trλ) and emission (Emλ) spectra:

Equation (6)

and

Equation (7)

Note that Trλ ≡ (Rp,λ/R*)2 and ${\mathrm{Em}}_{\lambda }\equiv {F}_{p,\lambda }/{F}_{* }$ in noiseless cases. These Trλ and Emλ signals were computed for the chosen system, instrument, and observation parameters for each planetary model in Table 1. We binned spectral channels to achieve a final resolution of R ≤ 100 per wavelength bin; instrument modes with higher intrinsic R were binned to R = 100, while ones with lower R (i.e., MIRI LRS at λ < 8 μm; see Kendrew et al. 2015) were not binned. We chose R = 100 as a compromise value low enough to maximize signal-to-noise while also being high enough to resolve molecular band spectral features.

We did perform a preliminary test of how binning impacts the retrieved uncertainties of H2O, CO, and CO2 in the cloudy transmission spectrum of a hot Jupiter of solar composition. We found that binning λ = 1–5 μm (NIRISS + NIRCam) spectra to R = 350 (two NIRISS resolution elements) did not produce any mixing ratio uncertainties that were less than ones retrieved for data binned to R = 100 in two trials using different instances of Nλ (no systematic noise was included). However, the ideal binning of each observing mode will likely be sensitive to actual in-flight noise performance, and this will not be known until after JWST begins operations.

The large aperture of JWST will ensure that the observatory will collect a large number of photons from bright stars, potentially resulting in the detection of ∼1010 or more photoelectrons per spectral bin when summing a significant number of observations of transit or eclipse events. In such cases, stellar photon noise will dominate Nλ, resulting in signal-to-noise ratios S/N ∼ 105 or Nλ values being only 10 parts per million (10 ppm) of signal values. Existing observations of transiting planets have not yielded such low noise values. Astrophysical noise (e.g., Barstow et al. 2015) and/or instrumental noise (e.g., decorrelation residuals) produce systematic noise floors that are not lowered when summing more data. The best HST WFC3 G141 observations of transiting systems to date have noise of the order of 30 ppm (Kreidberg et al. 2014a), while observations of transiting planets with the Spitzer space telescope Si:As detectors showed noise as low as ∼65 ppm (Knutson et al. 2009).

We adopt reasonably optimistic systematic noise floor values of 20, 30, and 50 ppm for NIRISS SOSS (λ = 1–2.5 μm), NIRCam grism (λ = 2.5–5.0 μm), and MIRI LRS (λ = 5.0–11 μm) observations, respectively. These are less than or equal to the values estimated by Deming et al. (2009) for the JWST NIRSpec and MIRI instruments. The excellent spatial sampling of the NIRISS GR700XD SOSS grism approaches that of the HST WFC3 G141 spatial scanning mode, and both instruments have reasonably similar HgCdTe detectors. We anticipate that decorrelation techniques will continue to improve, so we assign a 20 ppm noise floor value to NIRISS even though HST has not yet done quite this well. NIRCam will have similar detectors but will not be sampled as well (see Table 4), so we assign it a systematic noise floor equivalent to HST's best performance to date. The MIRI imager/LRS detector is similar (in materials, architecture, and sampling) to the Spitzer IRAC Si:As detector used by Knutson et al. (2009), and we use this as the basis for assigning a 50 ppm noise floor to MIRI. We will not know the actual performance of these instruments until after JWST commissioning, but Beichman et al. (2014) have demonstrated decorrelated noise precision similar to the adopted NIRISS and NIRCam values in laboratory tests devised to simulate their observations of transiting planets.

A single instance of noise was computed for each transmission or emission observation, using the expression for Nλ and propagating through the relation for Trλ or Emλ. We then added the appropriate systematic noise floor in quadrature, using the values given above. The single noise instance was drawn from a Poisson distribution with this amplitude, and the instance was added to the Trλ or Emλ signal to produce the final simulated spectrum for each planet model in Table 1. Uncertainties were set to the total noise amplitude determined for each bin; Figures 4 and 5 show the simulated spectra for each case with these uncertainties plotted as error bars for one example. We computed different noise instances for each instrument mode (NIRISS SOSS, NIRCam LW grism, and MIRI slitless LRS). The resultant spectra and uncertainties were then used to retrieve the parameters of each model planet's atmosphere as discussed in Section 3.

Figure 4.

Figure 4. Final simulated transmission spectra of the transmission models shown in Figure 2. The ${\mathrm{Tr}}_{\lambda }={({R}_{p,\lambda }/{R}_{* })}^{2}$ spectra are for a single transit with equal time on the star alone for each of the four instrument modes (Table 4). Spectra have been been binned to resolution R ≤ 100 (hot Jupiter, warm Neptune, warm sub-Neptune; see Section 4.2) as used for all retrievals or R = 35 (cool super-Earth) for display purposes only. The simulated spectra include a noise instance and are presented as colored curves. The black error bars denote 1σ of noise composed of random and systematic components. Dashed lines show the wavelength range boundaries of the chosen NIRISS, NIRCam, and MIRI instrument modes. The data used to create this figure are available.

Standard image High-resolution image
Figure 5.

Figure 5. Final simulated spectra of the emission models shown in Figure 3. The ${\mathrm{Em}}_{\lambda }=S({F}_{p})/S({F}_{* })$ spectra are for a single eclipse with equal time on the star alone for each of the four instrument modes (Table 4). Spectra have been been binned to resolution R ≤ 100 as used for all retrievals (hot Jupiter; see Section 4.2) or R = 35 for display purposes only (warm Neptune, warm sub-Neptune, cool super-Earth). The simulated spectra include a noise instance and are presented as colored curves. The black error bars denote 1σ of noise composed of random and systematic components. Dashed lines show the wavelength range boundaries of the chosen NIRISS, NIRCam, and MIRI instrument modes. Only one model is shown for the cool super-Earth for clarity; the noise is much greater than the difference between the clear solar and 100% H2O atmospheres. The data used to create this figure are available.

Standard image High-resolution image

5. RETRIEVAL RESULTS

The retrievals were performed for the wavelength ranges of up to three different instrument combinations for each of the model planet atmospheres: NIRISS (λ = 1.0–2.5 μm), NIRISS + NIRCam (λ = 1.0–5.0 μm), and NIRISS + NIRCam + MIRI LRS (λ = 1.0–11 μm). Retrievals were performed for all three combinations in cases of high S/N, i.e., the transmission spectra of all planet atmospheres and the hot-Jupiter emission spectra. The warm-Neptune and warm-sub-Neptune emission spectra had insufficient S/N for retrievals with only NIRISS (1.0–2.5 μm) data, and the complete λ = 1–11 μm emission spectrum for a cool super-Earth had insufficient S/N for retrievals. There simply is not enough flux contrast Fp/F* for useful emission spectra from this system when a single secondary eclipse is observed at each wavelength. Photometric filter observations may be more useful for constraining the planet's properties. Small (R ≲ 2 R), cool (T < 700 K) planets will need host stars with K ≲ 8.5 mag and/or spectral types later than M0 V for useful emission spectra of single secondary eclipses.

In this section, we focus on the retrieved parameters that most directly impact the simulated spectra: mixing ratios of significant molecular absorbers (CH4, CO, CO2, H2O, NH3), clouds, and atmospheric temperature–pressure (TP) profiles. C/O (carbon-to-oxygen ratios) and [Fe/H] are derived (not retrieved) from these quantities by a Monte Carlo propagation of the molecular uncertainties as in Line et al. (2013b) and Line et al. (2015), respectively.

Figures 69 summarize the marginalized posteriors for the relevant retrieved parameters and the derived C/O and [Fe/H] for the different planet and atmosphere scenarios. The true values of the input forward model and retrieval priors are also indicated in the figures. Any offsets between retrieved distribution medians and true values are due to the particular instance of random noise on the simulated spectra. Note that the input [Fe/H] values are slightly lower than the solar ([Fe/H] = 0) and 1000× solar ([Fe/H] = 3) bulk values that were used to construct the forward models (see Section 2). This is because some of the O atoms have been removed to account for the expected formation of Mg2SiO4 (enstatite) condensates in clouds in deep atmospheres below the regions probed by emission or transmission spectra (except in the hot Jupiter).14

Figure 6.

Figure 6. Retrieved and derived quantities of the hot Jupiter planet. Retrieved volume mixing ratios are shown for each molecular species. Marginalized posterior histogram shadings are color-coded by instrument mode and therefore wavelengths of spectra used for retrievals. Priors are indicated by blue horizontal dashed lines for the retrieved molecules. The resulting distributions for the derived quantities log(C/O) and [Fe/H] from the gas priors are more complex distribution functions (shown in blue). True values are indicated by vertical dashed black lines for all quantities.

Standard image High-resolution image
Figure 7.

Figure 7. Retrieved and derived quantities of the warm Neptune planet. Retrieved mixing ratios are shown for each molecular species. Marginalized posterior histogram shadings are color-coded by instrument mode and therefore wavelengths of spectra used for retrievals. Priors are indicated by blue horizontal dashed lines for the retrieved molecules. The resulting distributions for the derived quantities log(C/O) and [Fe/H] from the gas priors are more complex distribution functions (shown in blue). True values are indicated by vertical dashed black lines for all quantities.

Standard image High-resolution image
Figure 8.

Figure 8. Retrieved and derived quantities of the warm sub-Neptune planet. Retrieved mixing ratios are shown for each molecular species. Posterior predictive histogram shadings are color-coded by instrument mode and therefore wavelengths of spectra used for retrievals. Priors are indicated by blue horizontal dashed lines for the retrieved molecules. The resulting distributions for the derived quantities log(C/O) and [Fe/H] from the gas priors are more complex distribution functions (shown in blue). True values are indicated by vertical dashed black lines for all quantities.

Standard image High-resolution image
Figure 9.

Figure 9. Retrieved and derived quantities of the cool super-Earth planet. Retrieved mixing ratios are shown for each molecular species. Posterior predictive histogram shadings are color-coded by instrument mode and therefore wavelengths of spectra used for retrievals. Priors are indicated by blue horizontal dashed lines for the retrieved molecules. The resulting distributions for the derived quantities log(C/O) and [Fe/H] from the gas priors are more complex distribution functions (shown in blue). True values are indicated by vertical dashed black lines for all quantities.

Standard image High-resolution image

The prior distributions on the retrieved parameters are uniform-in-log. However, uniform-in-log priors bias the resulting C/O and [Fe/H] distributions (see discussion in Line et al. 2013b), thus we show what these distributions look like as a result of the uniform-in-log priors on the gas mixing ratios. These would be the distributions obtained had we made no observations. Therefore these prior-derived distributions must be considered in interpreting the posterior log(C/O) and [Fe/H] distributions.

5.1. Impacts of Atmospheric Parameterization

We now investigate whether the chosen atmospheric parameterization and priors may impact the retrieved results. Kreidberg et al. (2015) used both the "free" and "chemically consistent" retrieval approaches to determine the C/O in the transmission spectrum of WASP-12b. They found consistent results between the two approaches, suggesting a robust solution. We perform a similar analysis here. In addition to retrieving the mixing ratios of the individual molecules independent of the chemistry that links them (the "free" approach performed thus far), we also retrieve log(C/O) and [Fe/H] directly as a test case (with prior ranges from −2 to 2, and −4 to 4 respectively). These quantities, together with the temperature and pressure at each level in the atmosphere, permit the determination of the thermochemical equilibrium gas mixing ratios. These solutions are advantageous as they are chemically plausible and thermochemically self-consistent. However, such an approach does not account for the nearly limitless possible combinations of physical and chemical processes occurring in planetary atmospheres (e.g., vertical mixing, photochemistry, ion chemistry, 3D transport, cloud-gas microphysics interactions, interior-atmosphere chemistry coupling, escape processes, non-LTE, etc.) so we have not adopted it as our primary technique.

While not comprehensive, we perform one example using this chemically consistent approach and examine the impact it has on our ability to infer the atmospheric metallicity and C/O ratio. We retrieve on the full 1–11 μm wavelength emission spectrum of the warm sub-Neptune for the clear solar-composition atmosphere. Figure 10 illustrates the resulting marginalized log(C/O) and [Fe/H] metallicity posteriors compared with those derived from the molecular abundances in our baseline free retrievals (i.e., Figure 8, top row). The C/O histograms are qualitatively similar; both suggest a weak constraint of the C-to-O ratio. Perhaps more interesting is the comparison of the metallicity histograms. The chemically consistent approach provides a metallically constraint that is several orders of magnitude better than that derived from retrieving the molecular mixing ratios freely. This is because the chemically consistent approach rules out combinations of molecular abundances that do not abide by thermochemical equilibrium. Effectively, more prior information is being added to the chemically consistent retrieval system in the form of a more sophisticated parameterization with more assumptions but with fewer free parameters. More generally, it would be possible to apply chemically consistent models on all posterior "free" retrieval histograms (Figures 69) to rule out non-physical parameter spaces within the equilibrium framework. This would be an intermediate step between the classic retrieval and forced self-consistency, but we do not implement it here.

Figure 10.

Figure 10. Comparison between the classic "free" retrieval approach used in this work and the "chemically consistent" approach for the λ = 1–11 μm emission spectrum of the warm sub-Neptune solar-composition atmosphere. The gray solid histograms show the same values as in the top row of Figure 8 (1–11 μm wavelengths). The black unfilled histograms are from the thermochemically consistent retrieval. The log(C/O) histograms are in good agreement, both indicating poor constraint of C/O for this planetary atmosphere. However, the [Fe/H] metallicity histograms are substantially different due to the imposition of thermochemical consistency.

Standard image High-resolution image

Given a high enough signal-to-noise ratio and sufficient spectral resolution along with correct physics and chemistry constraints, one would expect the two approaches to produce the same distributions of the retrieved quantities. That would suggest true independence from any prior assumptions, and this would be the ideal regime for learning more about these atmospheres.

5.2. T–P Profiles and Parameter Uncertainties

Figure 11 shows the range of TP profiles retrieved from the simulated emission spectra over the three different wavelength ranges. Figure 12 shows normalized thermal emission contribution functions for the solar-composition hot Jupiter and warm sub-Neptune planets to illustrate where their thermal emissions originate. The contribution function for the solar-composition warm Neptune is very similar to that for the warm sub-Neptune.

Figure 11.

Figure 11. Temperature–Pressure (TP) profiles retrieved from emission spectra. True values are shown as dashed lines. Solid lines are best-fit retrieved values, and shaded regions denote 1σ uncertainties. Note that shadings are color-coded by instrument mode and therefore the wavelengths of spectra used for retrievals. Sections of the temperature–pressure profiles outside well constrained regions (see text and Figure 12) are extrapolated from the temperature profile parameterization.

Standard image High-resolution image
Figure 12.

Figure 12. Normalized thermal emission contribution functions for the solar-composition hot Jupiter (left) and warm sub-Neptune (right) planets. The figure shows the derivative of transmittance with altitude weighted by the Planck function at that level. Dark red areas are where the emission predominantly originates whereas blue areas represent little or no emission. The pressure levels probed by the red areas contribute most to the retrieved composition and thermal structure information (e.g., see Line et al. 2013b, 2014b). The contribution function for the warm Neptune of solar composition is very similar to that for the warm sub-Neptune.

Standard image High-resolution image

Table 5 lists the 68% confidence intervals of the retrieved parameters of the transmission and emission scenarios (see Section 3, Table 1). For cases in which a molecule is not particularly abundant (e.g., CH4 in the hot Jupiter and CO or CO2 in the cooler objects), only upper limits could be obtained. We quote the 3σ upper limits instead of confidence widths in these cases. We caution, however, that many of these upper limits are relatively soft and that one should really look at the histograms to get a sense of the distribution.

Table 5.  Retrieved Parameter Uncertainties or Limitsa

Planet λ (μm) Geom. T (K) xRp log(Pc) (bar) log H2O log CH4 log CO log CO2 log NH3 log N2
Hot Jupiter
Clear solar 1–11 Trans. 25 0.002 >−0.82 0.16 <−7.1 0.32 0.23 <−7.1 <−2.7
  1–5.0 Trans. 28 0.002 >−0.77 0.17 <−7.1 0.34 0.25 <−6.5 <−2.7
  1–2.5 Trans. 35 0.002 >−0.82 0.20 <−6.0 1.8 <−5.1 <−6.0 <−2.5
Cloud solar 1–11 Trans. 18 0.01 1.1 1.1 <−5.6 1.3 2.9b <−5.5 <−2.4
  1–5.0 Trans. 33 0.011 1.3 1.5 <−5.4 1.6 3.3b <−4.6 <−1.8
  1–2.5 Trans. 180 0.021 1.6 1.5 <−3.3 <−1.3 <−2.6 <−3.6 <−0.80
Solar 1–11 Emis. 0.17 <−6.9 0.26 0.17 <−6.4
  1–5.0 Emis. 0.22 <−6.9 0.36 0.23 <−5.9
  1–2.5 Emis. 0.63 <−4.6 <−1.1 <−4.6 <−5.1
Solar 1–11 Emis. 0.32 <−6.1 0.38 0.41 <−5.5
inversion 1–5.0 Emis. 0.36 <−5.8 0.44 0.53 <−5.2
  1–2.5 Emis. 3.6 <−2.5 <−2.5 <−2.5 <−5.0
Warm Neptune
Clear solar 1–11 Trans. 17 0.008 >−1.3 0.59 0.51 <−3.5 <−7.1 0.53 <−2.5
  1–5.0 Trans. 19 0.010 >−1.4 0.71 0.61 <−3.3 <−7.1 0.63 <−2.4
  1–2.5 Trans. 42 0.014 >−1.7 0.91 0.87 <−2.3 <−3.7 0.85 <−2.3
Cloud solar 1–11 Trans. 76 0.025 1.6 1.7 1.7 <−3.1 <−5.0 1.7 <−1.2
  1–5.0 Trans. 74 0.020 1.5 1.6 1.5 <−2.9 <−4.9 3.6b <−1.7
  1–2.5 Trans. 180 0.029 1.9 2.0 1.9 <−0.9 <−2.4 <−1.8 <−0.50
High MMW 1–11 Trans. 170 0.005 >−3.2 3.9b 1.6 <−0.10 1.6 <−2.5 <−0.20
  1–5.0 Trans. 240 0.004 >−2.8 3.7b 1.8 <−0.20 1.8 <−2.7 <−0.20
  1–2.5 Trans. 300 0.005 >−3.4 <−0.06 >−4.0 <−0.20 2.7 <−1.9 <−0.20
Solar 1–11 Emis. 1.26 0.79 <−3.6 <−6.1 1.2
  1–5.0 Emis. 2.4 1.5 <−2.4 <−5.7 4.7b
High MMW 1–11 Emis. <−1.6 2.1 <−1.2 >−6.2 <4.1
  1–5.0 Emis. 4.3 <−1.8 <−0.7  
Warm Sub-Neptune
Clear solar 1–11 Trans. 8.1 0.006 >−0.70 0.25 0.19 <−6.4 <−8.0 0.21 <−2.0
  1–5.0 Trans. 9.8 0.009 >−0.8 0.40 0.30 <−6.1 <−7.9 0.31 <−2.5
  1–2.5 Trans. 33 0.029 >−1.6 1.1 1.0 <−2.4 <−3.7 1.0 <−2.1
Cloud solar 1–11 Trans. 22 0.024 0.92 1.0 0.96 <−4.0 <−5.7 1.0 <−2.4
  1–5.0 Trans. 25 0.044 1.6 1.7 1.6 <−5.4 <−5.9 1.6 <−2.2
  1–2.5 Trans. 170 0.076 1.9 2.1 2.1 <−1.2 <−6.4 2.4 <−0.80
High MMW 1–11 Trans. 100 0.004 >−3.3 1.0 0.65 <−0.4 0.84 <−2.4 <−0.10
  1–5.0 Trans. 130 0.005 >−3.2 1.4 1.1 <−0.10 1.4 <−2.6 <−0.10
  1–2.5 Trans. 200 0.007 >−3.4 >−5.0 1.2 <−0.20 <−0.50 <−1.8 <−0.10
Solar 1–11 Emis. 2.2 1.5 <−1.4 <−4.7 2.2
  1–5.0 Emis. 6.1b 2.5 <−1.7 <−4.6 <−0.50
High MMW 1–11 Emis. 4.3b 2.3 <−0.50 3.2b <−0.80
  1–5.0 Emis. <−0.10 <−2.1 5.0b
Cool Super-Earth
Clear solar 1–11 Trans. 72 0.030 >−2.4 1.8 1.3 <−2.1 <−3.4 1.3 <−1.1
  1–5.0 Trans. 85 0.028 >−2.5 1.9 1.3 <−2.1 <−3.3 1.3 <−1.1
  1–2.5 Trans. 120 0.035 >−2.8 <−0.7 1.4 <−1.4 <−1.9 1.4 <−0.20
Cloud solar 1–11 Trans. 570 0.25 2.8 <−0.10 >−7.5 <−0.11 <−0.44
  1–5.0 Trans. 700 0.200 >1.5 <−0.10 >−7.5 <−0.28 <−0.27
  1–2.5 Trans. <3000 0.33 >1.5
Clear H2O 1–11 Trans. 670 0.043 >−5.9 >−5.0 <−1.8 <−0.10 <−0.14
  1–5.0 Trans. 720 0.044 >−5.9 >−7.0 <−1.4 <−0.10 <−0.50
  1–2.5 Trans. 970 0.075 >−5.9 >−7.0 <−0.85 <−0.10 <−0.36

Note. 68% confidence widths are given when parameters are well determined, otherwise 3σ upper (<) or lower (>) value limits are listed. No constraint (the probability spans the prior range) is denoted by "–" non-retrieved values are denoted by "⋯." To aid interpretation, an uncertainty of 0.3 corresponds to a constraint within a factor of 2, 0.7 within a factor of 5, and 1 within a factor of 10 for the log value parameters. All values are uncertain to ∼10% due to statistical jitter in MCMC retrievals and in the random noise instance applied to the simulated data.

aRetrieved parameters for transmission (Trans.) and emission (Emis.) spectra are described in Section 3. b68% confidence width, but lower mixing ratio values are unconstrained at 2σ.

Download table as:  ASCIITypeset image

These numbers are meant to be a guide to illustrate how the constraints change from one object to the next. We also note that there is typically a ∼10% uncertainty on these uncertainty values due to the particular instances of random noise applied. In the remainder of this section, we examine how well the directly retrieved parameters constrain the different planetary atmospheres.

5.3. Constraints from Transmission Spectra

Figures 69 and Table 5 show that the mixing ratios of the dominant molecules for a particular planet type are well constrained (bounded posterior rather than an upper limit) in the transmission spectra of all clear solar atmospheres in most cases. Furthermore, observations with NIRISS alone (λ ≤ 2.5 μm) can constrain H2O in these atmospheres nearly as well as observations out to 11 μm wavelengths in some cases. For example, Table 5 shows that the retrieved H2O mixing ratio of the hot Jupiter with a clear solar atmosphere has a 68% uncertainty of 0.16 dex (45%) for a λ = 1–11 μm spectrum, and this degrades to only 0.20 dex (58%) when only using the 1–2.5 μm wavelengths in a NIRISS observation. However, CO is well constrained (uncertainty ∼0.3 dex for retrievals done with the full 1–11 μm or 1–5 μm spectra), but the uncertainty is much worse (∼1.8 dex) when only the NIRISS data are used.

The H2O and CH4 uncertainties of the warm Neptune with clear solar atmosphere increase by ∼0.3 dex (factor of ∼2), providing uncertainties of ∼1 dex (factor of ∼10) when using only the NIRISS wavelengths. These species have similar NIRISS-only uncertainties in the warm sub-Neptune with clear solar atmosphere, but their full λ = 1–11 μm spectra provide better uncertainties (∼0.2 versus ∼0.5 dex for the warm Neptune). These values are about ∼1.5 dex for the full 1–11 μm wavelengths for the cool super-Earth with clear solar atmosphere, and the CH4 uncertainty does not change much when considering only the NIRISS data range. However, the retrieved H2O mixing ratio value becomes an upper limit. Overall it is encouraging that these atmospheres can be constrained so well in these modest observations (one transit at each wavelength), and the wavelength range of NIRISS alone is sufficient to detect dominant species with good confidence or even measure mixing ratios to good precision in some cases (i.e., the chosen hot Jupiter). Even for gases that are not abundant in the atmospheres (CH4 and NH3 in the hot Jupiter, CO and CO2 in the cooler objects), relatively low upper limits can be obtained. For instance, most scenarios within the hot Jupiter planet type can detect, or rather rule out, NH3 and CH4 at the ∼1 ppm level.

Planets with cloudy atmospheres will be more challenging to constrain with transmission spectra; this is already apparent from numerous HST observations of GJ 1214b (e.g., Croll et al. 2011; Kreidberg et al. 2014a, and references therein) and other planets that exhibit weak or no spectral features. Figures 69 and Table 5 show that the mixing ratios of most molecules are not constrained well (uncertain by more than 1.0 dex) with λ ≤ 2.5 μm (NIRISS only) transmission spectra for exoplanet atmospheres of solar composition with clouds. Mixing ratio uncertainties improve somewhat when λ ≥ 5 μm data (NIRCam or NIRCam+MIRI) are added in many cases. However, the resulting constraints are considerably worse than those of the clear solar atmospheres (as good as 0.2–0.5 dex as discussed above). Mixing ratio uncertainties improve to ∼1 dex for some molecules in the cloudy atmospheres of the hot Jupiter and warm sub-Neptune when these longer wavelengths are also included in the retrievals. The mixing ratio uncertainties of the molecules in the warm Neptune system also improve with these longer wavelength retrievals, but none gets close to 1.0 dex. The cool super-Earth system retrievals only produce molecular mixing ratio limits for the cloudy atmosphere case regardless of wavelength range used.

Cloud-top pressure Pcloud is also retrieved for the transmission cases, and Table 5 shows that this is constrained to ∼1 dex for the hot Jupiter and warm sub-Neptune systems when using complete λ = 1–11 μm spectra of cloudy solar atmospheres. This allows the position of clouds to be located to an order of magnitude in pressure in their atmospheres. Reducing the wavelength range to 1–5 and 1–2.5 μm increases the uncertainty to ∼1.4 and ∼1.7 dex, respectively. The warm Neptune system is not constrained quite as well (mostly due to its less favorable transmission parameters), having an uncertainty in Pcloud of 1.4–1.9 dex for the three different wavelength ranges of the cloudy solar atmosphere case. The retrievals provide lower limits for Pcloud (highest cloud altitudes) for all clear atmospheres (solar or HMMW) of the hot Jupiter, warm Neptune, and warm sub-Neptune systems. Clouds in the clear solar atmospheres of all three systems are constrained to P ≳ 1 mbar, a useful limit given that transmission spectra probe only high-altitude/low-pressure atmospheric regions. Pcloud is not constrained well for any atmospheres of the cool super-Earth system. The Pcloud histograms in Figures 69 illustrate these constraints graphically.

5.4. Constraints from Emission Spectra

Emission spectra retrievals constrain the mixing ratios of the most dominant species of the solar composition atmospheres of the hot Jupiter (better than 0.3 dex for CO, CO2, and H2O) and warm Neptune (to ∼1 dex for CH4, H2O, and NH3). Figures 69 and Table 5 show that λ > 5 μm data (e.g., MIRI) are required to obtain these good constraints. This can be understood by examining Figure 5; emission spectra have low S/N at shorter wavelengths, and there are numerous strong molecular absorption features at λ > 2.5 μm. Note that the NIRISS-only (λ = 1–2.5 μm) emission spectrum of the hot Jupiter gives a false peak in its CH4 mixing ratio, and this disappears when longer wavelength data are added (see Figure 6). Molecular mixing ratio uncertainties are worse than ∼1 dex for the emission retrievals of the warm sub-Neptune planet, and the S/N of the emission spectrum of the cool super-Earth was too low to perform useful retrievals at any wavelengths (see also Figure 5).

These results show that JWST λ > 2.5 μm emission spectra with moderate-to-high S/N will be very useful for atmospheric characterization. Emission spectra are also required to retrieve TP profiles (Figure 11), which are particularly important for understanding energy absorption and transport in strongly insolated atmospheres. Fortney (2005) showed that the direct, face-on geometry relevant to emission spectra is much less impacted by clouds or hazes than the slant geometry of transmission spectroscopy observations. Therefore we assume that observations and retrievals of emission spectra from a cloudy solar atmosphere will be very similar to those from a clear solar-composition atmosphere studied here. Given that, retrievals of JWST λ = 1–11 μm emission spectra provide significantly better constraints than the transmission spectra on the mixing ratios of most molecules for cloudy hot-Jupiter and warm-Neptune atmospheres (see Table 5). Each star+planet system should be evaluated to determine whether emission or transmission observations are more favorable for detecting spectral features and constraining parameters of interest via atmospheric retrievals. It would be ideal to acquire and combine both transmission and emission spectra to probe planetary atmospheres over the broadest possible pressure range and to use all data to constrain compositions and chemistries (Griffith 2014; Kreidberg et al. 2014b).

Spectral absorption features are clearly suppressed over the entire λ = 1–11 μm wavelength range of the hot-Jupiter emission spectrum with a temperature inversion (Figure 5), and the inversion is detected at modest to high S/N in the retrievals. As shown in Figure 11, the temperature of the hot-Jupiter atmosphere decreases to T = 1260 K at P = 10−2 bar, and then increases to T = 2100 K at P = 10−4 bar for the inversion model. This difference of ΔT = 840 K is detectable in the retrieved TP profiles of all three wavelength ranges. At these pressures, the 1σ temperature uncertainties of the 1–2.5, 1–5, and 1–11 μm wavelength ranges are approximately 215 K, 40 K, and 30 K respectively. Therefore the temperature inversion is detected at better than 4σ, 21σ, and 28σ respectively when using multiple retrieved TP points in this pressure range. It is clear that the λ > 2.5 μm data constrain the hot-Jupiter TP profile much better than the λ = 1–2.5 μm data alone, and Figure 11 shows that λ = 1–11 μm data constrain TP profiles somewhat better than λ = 1–5 μm data for all three planets studied.

5.5. HMMW Atmospheres and [Fe/H]

It will also be possible to constrain the compositions of HMMW (including 100% H2O) atmospheres with JWST spectra. Figures 79 show that the transmission retrievals of the HMMW atmospheres of the warm Neptune, warm sub-Neptune, and cool super-Earth have significantly different molecular mixing ratio distributions than the solar ones (with or without clouds; as expected from Table 3). Likewise, their [Fe/H] distributions are also significantly different, reflective of the high metallicity ([Fe/H] ≃ 3) required to produce HMMW atmospheres. These differences are readily apparent in the λ = 1–2.5 μm results alone in most cases, but they become clearer when longer wavelength data are included. Figures 78 and Table 5 clearly show that the emission retrievals constrain the clear HMMW atmospheres of the warm Neptune and warm sub-Neptune less well than transmission ones. These figures also show that the derived [Fe/H] values have uncertainties of ∼0.5 dex (a factor of 3) or better for the clear (HMMW and solar composition) atmospheres of the hot and warm planets studied with λ = 1–5+ μm transmission spectra.

5.6. Potential Missing Species and Opacity Uncertainties

Real planets may have atmospheric species not included in the forward models or retrievals we have performed here. If this were the case, performing these retrievals on real JWST data would produce wavelength-dependent residual differences between observed and modeled spectra. Additional species could then be introduced into the model to minimize these residuals. Computing the Bayes factor between the models with simpler and more complex compositions would reveal whether adding the additional species is statistically justified. Uncertain molecular opacities may also skew the retrieved compositions for observed planets. However, this error is likely to be of the order of 10% but can be as high as a factor of a few in some cases (Grimm & Heng 2015, Section 3.4). This is mostly less than or equal to our best 68% mixing ratio confidence widths (see Table 5).

6. DISCUSSION

Here we discuss how applying these precision molecular abundance and temperature constraints may address several significant outstanding scientific issues. We focus on questions probing planet formation and disequilibrium chemistry and end with observational considerations that will impact data quality and the uncertainties of retrieved values.

6.1. Carbon-to-oxygen Ratios

The C-to-O ratio (C/O) of a planetary atmosphere could be a tracer of its formation and migration history within a protoplanetary disk (Öberg et al. 2011). Determining the C/Os of a diverse set of exoplanets over a wide range of conditions is important for applying this theory to understand planet formation scenarios. There has yet to be unambiguous evidence for high C/O (C/O > 1) within the current ensemble of observed exoplanets (Madhusudhan et al. 2011; Konopacky et al. 2013; Line et al. 2014b; Stevenson et al. 2014; Barman et al. 2015; Benneke 2015; Kreidberg et al. 2015). Only upper limits (Benneke 2015; Kreidberg et al. 2015) have been derived from transmission spectra. Emission spectra retrievals have also provided upper limits for some planets that are otherwise unconstrained, and a specific C/O value has only been determined for HD189733b, which has broad near-continuous wavelength coverage from ∼1–20 μm (Lee et al. 2012; Line et al. 2014b; Waldmann et al. 2015).

There are several ways of determining or diagnosing the C/O in exoplanet atmospheres. The most straightforward, direct approach is to simply determine the abundance of all of the carbon- and oxygen-bearing species present in the spectra and compute it directly by summing the carbon atoms in all of the carbon-bearing species and dividing by the oxygen atoms in all the oxygen-bearing species. Certainly it is possible for some carbon or oxygen to be locked away in condensates or other species that may not necessarily present themselves spectroscopically, resulting in a potential bias. In any case, this is the approach we take. Figures 69 show the C/O histograms derived from the retrieved molecular abundances. In some cases, when the abundances are not well constrained, a two-peak distribution exists. This is simply due to the propagation of uniform-in-log priors through the ratio used to compute the C/O (see Line et al. 2013b, for an in-depth discussion). Another approach is to retrieve the C/O directly as done by Kreidberg et al. (2015), Benneke (2015), and in Section 5.1 for a test case. However, doing so requires the assumption of the chemical processes (e.g., thermochemical equilibrium, vertical mixing, photochemistry) via a chemical model that relates the elemental abundances to the molecular abundances (see Section 5.1).

We find that for all of the hot-Jupiter transmission and emission scenarios observed with wavelengths longer than 2.5 μm (e.g., NIRISS+NIRCam and NIRISS+NIRCam+MIRI) the C/O can be constrained to better than 0.2 dex (a factor of 1.6) and is largely unbiased by the two-peak problem. Wavelengths below 2.5 μm do not capture the strong CH4, CO, or CO2 vibrational bands that occur at longer wavelengths. The low abundance of CH4 requires observation of the strong 3.3 and/or 7.7 μm band to provide a meaningful upper limit as CH4 is not thermochemically dominant in hot atmospheres. While CO and CO2 are dominant in hot atmospheres, they have a relatively narrow influence across wavelength. Furthermore, there is little improvement in extending coverage beyond 5 μm as the strongest CO or CO2 bands over the full 1–11 μm wavelength range occur at ∼4.5 μm (see Figure 1). Therefore, for hot Jupiters with comparable S/Ns, observations over λ = 1–5 μm are more than sufficient to provide a meaningful constraint on their C/O.

The transmission spectra of clear atmospheres of low mean molecular weight (solar composition) on the warm Neptune and sub-Neptune also unbiasedly constrain the C/O to ∼0.2 dex or better in all three wavelength ranges. This is because the CH4 is largely constrained to abundances that are greater than the CO or CO2 abundances and thus the C/O ratio is mainly set by the methane-to-water ratio (CH4/H2O). Because of the large abundance of CH4 in the cooler planets, it presents strong spectral features at wavelengths below 2.5 μm (see Figure 1). Even in the cloudy scenarios the log(C/O) constraints are only a factor of 1.2 worse.

Emission spectra constrain C/O less well in these small warm planets. 1–11 μm emission spectra constrain C/O to ∼0.5 dex in the solar-composition atmosphere of the warm Neptune, and 1–5 μm emission spectra also provide some constraint (to only ∼1 dex). The λ ≥ 5 μm emission spectra provide only loose constraints on C/O in the clear solar atmosphere of the warm sub-Neptune. Even complete 1–11 μm emission spectra produce C/O distributions very similar to the priors for the HMMW atmospheres of these planets. These relatively cool planets have low flux and therefore low S/N near the λ ≤ 4 μm CH4 bands, so these results are not particularly surprising.

In general, the C/O constraints are poor for the cool super-Earth atmospheres. None of the transmission scenarios constrain C/O to significantly better than an order of magnitude.

Finally, Figure 13 shows how the retrieved H2O mixing ratio uncertainties compare to the predicted thermochemical H2O values for two different C/O cases (see Kreidberg et al. 2015). The uncertainties in H2O mixing ratios are shown for the different planets' transmission cases (their positions are located for clarity and are not indicative of true temperatures or compositions). The H2O mixing ratio depends on metallicity and contains no direct information on carbon abundance, so it cannot provide an unambiguous C/O constraint by itself. However, it is instructive to show how one could use H2O to rule out a high C/O scenario in some cases. For hot Jupiters, where the difference between solar and high C/O chemical models is largest, a measurement in a clear atmosphere could distinguish the difference between a solar and high C/O atmosphere by greater than 50σ for all three wavelength ranges. Using H2O alone becomes difficult for planets with scale height temperatures below T ∼ 1000 K as the predicted thermochemical H2O abundances converge for both solar value and high C/O. Other proxies such as the ratio of H2O to CH4 (under the assumption that oxidized carbon will be in low abundance; see Figure 2 of Madhusudhan 2012) or a direct determination of the C/O (as above) will be required.

Figure 13.

Figure 13. Application of precision measurements of H2O and scale height temperature from transmission spectra to distinguishing between high and low C/O scenarios (after Kreidberg et al. 2015). The broad red and blue curves are equilibrium chemistry models that show how the H2O abundance changes as a function of temperature for high (red) and low (blue) C-to-O ratios for solar metallicity. The spread in the equilibrium chemistry models is due to different assumptions regarding the probed pressure levels (0.1–10 mbar). We show representative error bars derived from our temperature and H2O retrieval results for the solar metallicity scenarios observed in transmission (Table 5). The red error bars represent the NIRISS only, blue are for NIRISS+NIRCam, and black are for NIRISS+NIRCam+MIRI. The top set of error symbols show the constraints in the clear atmospheres and the bottom ones show cloudy ones. These positions are located for clarity and are not indicative of actual retrieved temperatures or compositions! The light gray points and error bars are constraints from HST WFC3 transmission measurements of WASP-12b (Kreidberg et al. 2015) (hotter) and WASP-43b KBD14b (cooler).

Standard image High-resolution image

Regardless of the methodology used, it appears that JWST spectra will be very useful in determining the C-to-O ratios in a wide array of planetary atmospheres and will allow us to begin to address quantitatively the formation location relative to ice lines and any subsequent migration.

6.2. The Mass–Metallicity Relationship

Understanding planet formation requires determining and interpreting the mass–metallicity relationship for a varied sample of planets. Establishing this relationship over the exoplanet population provides insight into the formation mechanism of core accretion (e.g., Ida & Lin 2005). There has been some progress in establishing the role that planetary mass and parent star metallicity have in determining the bulk metallicity of transiting gas giants (Miller & Fortney 2011) through the use of planetary evolution models. A corresponding constraint on planetary atmospheric metallicity, from spectroscopy, would help us to understand whether most of these metals are found in a core or are mixed within the H/He-dominated envelope.

Using population synthesis models (Mordasini et al. 2012), Fortney et al. (2013) suggest that as the mass of a planet decreases, the atmospheric metallicity increases. Lower mass planets are unable to accrete substantial envelopes within the core-accretion theory and thus are more susceptible to pollution by infalling planetesimals. We do indeed find tantalizing evidence for such a trend within our own solar system as seen in Figure 14. However, in our own solar system we are limited to using CH4 as the proxy for metallicity as water is largely sequestered in deep clouds. Exoplanet atmospheres, due to their high temperatures, permit us to access a wide array of molecules, allowing for more precise constraints on the envelope metallicity. Kreidberg et al. (2014b) provided additional leverage on this relationship via a relatively good constraint on the water abundance in a 2 MJ hot Jupiter. Under the assumption of solar C/O, the retrieved water abundance was used as a proxy for metallicity. This proxy-metallicity was found to be consistent with the observed solar system trend. Furthermore, constraints on Neptune-mass objects (GJ 436b by Stevenson et al. 2010 and HAT-P-11b by Fraine et al. 2014) are also suggestive of this trend.

Figure 14.

Figure 14. Atmospheric mass–metallicity relationship (after Kreidberg et al. 2014b). Solar system planets and a measured exoplanet are shown as the black points with error bars. Representative metallicity constraints from transmission spectra for the clear hot Jupiter of solar composition and the clear warm Neptune of high mean molecular weight are shown near the top (not at their actual mass/metallicity values). Three 68% confidence constraints are shown for each: black for NIRISS+NIRCam+MIRI, blue for NIRISS+NIRCam, and red for NIRISS only.

Standard image High-resolution image

Recently, Benneke (2015) demonstrated that the water abundance alone cannot constrain the atmospheric metallicity because of the degeneracy of metallicity with the carbon-to-oxygen ratio. The broad wavelength coverage and high S/N of JWST data enable us to constrain not only the water abundance, but carbon species as well. This, in essence, breaks the C/O–metallicity degeneracy. We directly compute the metallicity ([Fe/H]) from the retrieved molecular mixing ratios, and the metallicity histograms are shown in Figures 69 (see Section 5). Figure 14 compares a typical metallicity constraint for the hot Jupiter of solar composition and the warm Neptune of HMMW. Observing just five planets spaced logarithmically between a few Jupiter masses and a Neptune mass with such constraints (0.4 dex) would allow us to determine the mass–metallicity slope (in log space) with a 1σ uncertainty of 0.13.

6.3. Disequilibrium Chemistry

Disequilibrium processes are likely to play a role in sculpting the molecular abundances in exoplanet atmospheres (Liang et al. 2003; Zahnle et al. 2009; Line et al. 2010, 2011; Moses et al. 2011; Venot et al. 2012; Hu et al. 2013; Agúndez et al. 2014). These processes come in a variety of flavors including, but not limited to, vertical and horizontal mixing (Prinn & Barshay 1977; Cooper & Showman 2006), photochemistry (Yung & Demore 1999), ion chemistry (Lavvas et al. 2008), and biology (e.g., Holland 1994). The predicted dominant disequilibrium process in Jovian-type planets is due to vertical mixing and thus we focus on the ability of JWST to infer disequilibrium due to vertical mixing. Vertical mixing tends to set (or quench) the upper atmospheric abundance of certain molecules to a particular deep atmospheric value, which can result in an orders-of-magnitude enhancement (or depletion) relative to the expected equilibrium abundance at a higher altitude in the atmosphere. Line & Yung (2013) devised a scheme to determine the degree to which an atmosphere is out of equilibrium given the retrieved abundances of H2O, CH4, CO, and H2. If these values form a ratio, α,

Equation (8)

where fi are the molecular mixing ratios of species i, P is the pressure in bar, and T the temperature in K, which is equivalent to the thermochemical equilibrium constant Keq(T), then those species are considered to be in chemical equilibrium and thus disequilibrium mechanisms are weak or non-existent.

Figure 15 summarizes the constraints on α as provided by the different wavelength regions for emission spectra. We also note, as in Line & Yung (2013), that there could be up to a factor of ∼100 more uncertainty due to the spread in the probed pressure levels. Furthermore, in some cases only an upper limit on one of the three retrieved gases can be obtained. In these cases α would also only have an upper or lower limit. We find that the hot-Jupiter scenarios will have the best constraint on α. Unfortunately hot Jupiters are generally predicted to be in equilibrium (Line & Yung 2013), so the small error bar is not very useful for identifying disequilibrium in those planets. The objects cooler than ∼1000 K will likely show signs of disequilibrium due to the dredging up of CO (and also N2 in the NH3–N2 chemical system). This will result in an α that falls below the equilibrium line. The maximum expected deviation due to strong vertical mixing is only ∼2σ larger than the smallest warm-Neptune error bar, making the definitive detection of disequilibrium due to vertical mixing difficult. Perhaps the best way of approaching this problem is to determine α over a wide range of planetary effective temperatures to identify where deviations from equilibrium begin to occur. This transition is likely to occur between 1000 and 1200 K (e.g., Moses et al. 2011). This temperature region is also a good place to observe because the CH4 and CO abundances begin to thermochemically trade places. Their nearly equal abundances and the higher temperatures (relative to the warm Neptunes) will likely allow bounded constraints on both the CO and CH4 abundances, providing an actual bounded constraint on α.

Figure 15.

Figure 15. Diagnosing disequilbirum chemistry with JWST (adapted from Line & Yung 2013). The green line is the equilibrium constant as a function of temperature. Given the equilibrium molecular abundances of H2O, CH4, CO, and H2 at a given temperature and pressure, the quantity α (see Equation (8)) will fall on this line. If there are strong disequilibrium processes, α will deviate from the green line. α computed from the retrieved abundances for H2O, CH4, CO, and H2 for a variety of planets observed with HST and Spitzer is shown in light gray (Line & Yung 2013; Line et al. 2014b). We show representative error bars for α derived from the retrieved mixing ratios of emission spectra for the solar-composition emission scenarios of a hot Jupiter, warm Neptune, and warm Sub-Neptune for NIRISS only (red), NIRISS+NIRCam (blue), and NIRISS+NIRCam+MIRI (black). Their positions in the plot are arbitrary. Model uncertainty due to the uncertainty in the pressure levels probed (e.g., the width of the thermal emission contribution function across wavelength) is shown as the black error bar. The maximum expected deviation due to vertical mixing is shown as the green error bar. Finally, the prior uncertainty is shown as the red curve.

Standard image High-resolution image

We have broadly explored the detectability of disequilibrium due to vertical mixing. Certainly, photochemistry can produce species not included in this investigation (e.g., CmHn, HCN), especially in cooler planets in which CH4 is readily available for photolysis (Line et al. 2011; Miller-Ricci Kempton et al. 2012; Moses et al. 2013; Hu & Seager 2014; Venot et al. 2014). Shabram et al. (2011) determined that some of these photoproducts are potentially detectable in JWST observations of planets similar to the warm Neptune (GJ 436b) we have considered here.

6.4. Additional Observational Considerations

The precision of the retrieved parameters is limited by the particular star–planet systems chosen as well as by the planets themselves. For example, the H2O mixing ratio uncertainty using all wavelengths is considerably better for the clear solar atmosphere of the warm sub-Neptune (0.25 dex) versus the warm Neptune (0.59 dex) in transmission. These two planets have similar equilibrium temperatures and scale heights (see Table 2), so this difference is likely due to the higher S/N of the simulations for the warm sub-Neptune caused by the relatively high brightness and small size of the adopted host star, GJ 1214. Each star+planet system should be evaluated to determine whether emission or transmission observations are more favorable for detecting spectral features and constraining parameters of interest. The planetary system parameters given in Table 2 should be useful for scaling these results to other systems.

We do not know how well co-adding observations of multiple transits or secondary eclipses will improve our results at this time. The simulated single-transit and single-eclipse observations of our selected systems typically have total noise values only 10%–50% larger than our adopted noise floors (Section 4), so systematic noise assumptions have already significantly influenced the precision of our simulated data and retrieved information. Given this, co-adding more data would not substantially improve the results for these very observationally favored systems with bright host stars. We believe that our systematic noise assumptions are reasonable, but we will not know their exact values in different conditions (e.g., co-adding multiple spectra versus binning to lower resolution) until after JWST becomes operational.

Other instrumental and astrophysical noise sources may also impact JWST data. Barstow et al. (2015) considered the impact of systematic wavelength-invariant offsets between the spectra acquired in different JWST instrument modes. They found that the impact on retrieved parameters was minimal if even small regions of overlap between the modes were used to correct such errors to below the noise of their simulations. This is also likely to apply in our study if residual offsets between the different spectral regions are (corrected to) less than their respective adopted systematic noise floors. We note that the different instrument modes have more spectral overlap than the regions we selected in Table 4.

Barstow et al. (2015) also note that star spots can impact transmission (but not emission) spectra considerably (see also Pont et al. 2013) and they assess their impact on retrievals of planet parameters for both solar- and M-type (Teff = 3000 K) host stars. They find that the quality of retrieved information is not seriously degraded for a hot Jupiter with a Sun-like star that has a relatively high spot fraction (3% of its area). However, Barstow et al. (2015) find that the retrieved H2O mixing ratio can be up to an order of magnitude too high for the transmission spectrum of a hot Neptune planet with an M-type host star with a spot coverage of 10% if the planet does not occult any spots during transit. This error is considerably larger than the H2O mixing ratio uncertainty we expect for the warm Neptune with a clear solar-composition atmosphere and full 1–11 μm wavelength coverage (Table 5), and we share the concern with Barstow et al. (2015) that stellar activity could be the limiting factor for accurate retrievals of JWST observations of active stars. Simultaneous photometric observations at relatively short wavelengths (e.g., see Fraine et al. 2014) could be used to correct this effect, and this may be possible with NIRCam λ ≤ 2.4 μm imaging during spectral observations with the NIRCam λ = 2.5–5.0 μm grism.

7. SUMMARY AND CONCLUSIONS

We have generated forward models of transmission and emission spectra of generic hot Jupiter, warm Neptune, warm sub-Neptune, and cool super-Earth planets using the parameters of well known systems. We then performed simulations of slitless JWST observations using the instrument modes NIRISS SOSS over 1–2.5 μm, the NIRCam grisms over 2.5–5.0 μm (two exposures required), and the MIRI LRS over 5.0–11 μm. The NIRSpec instrument may be used instead of NIRCam, but it also requires two exposures over 2.5–5.0 μm for bright stars, and we understand its likely systematic noise less well than NIRCam at this time. We performed retrievals on single simulated transits and secondary eclipses to assess what information and constraints JWST data are likely to provide on these planet archetypes and other exoplanet atmospheres. We arrive at the following major conclusions:

  • 1.  
    JWST will likely obtain high-quality transmission and emission spectra of a variety of exoplanet atmospheres over a wavelength range of at least 1–11 μm. Obtaining spectra over this entire range will typically require observations of four separate transit or secondary eclipse events using four instrument modes for planets with bright host stars. We do not know at this time exactly how JWST data will be impacted by systematic noise, but we have assumed noise floors consistent with the best performance of HST or Spitzer as appropriate for each wavelength range.
  • 2.  
    The volume mixing ratios of dominant molecular species, C/O, and [Fe/H] of clear planetary atmospheres with solar composition can be diagnosed very well with transmission spectra, often with only 1–2.5 μm wavelength NIRISS data. However, longer wavelength data are needed for good constraints in some cases. Strong temperature inversions should be detectable in the TP profile of the clear solar-composition hot Jupiter planet with emission spectra covering λ = 1–2.5 μm. The TP  profiles of the hot Jupiter, warm Neptune, and warm sub-Neptune are all constrained reasonably well with 1–5 μm emission spectra. JWST transmission spectra will be useful for discriminating between cloudy and HMMW atmospheres of small planets in observationally favorable systems.
  • 3.  
    JWST spectra have the potential to constrain the molecular mixing ratios, C/O, [Fe/H], and TP profiles of planets with cloudy atmospheres. Mixing ratios of the cloudy warm sub-Neptune are constrained to ∼1 dex with transmission spectra covering λ = 1–11 μm. However, λ = 1–11 μm emission spectra provide significantly better constraints than the transmission spectra on the mixing ratios of most molecules for cloudy atmospheres of the hot Jupiter and warm Neptune, assuming that emission spectra are not impacted by clouds. Each star+planet system should be evaluated to determine whether emission or transmission observations are more favorable for detecting spectral features and constraining parameters of interest. Emission spectra can be more useful than transmission in cases of sufficiently high Fp accompanied by high Fp/F*. Acquiring both transmission and emission spectra will probe exoplanet atmospheres over the broadest possible pressure ranges and constrain compositions and chemistries better than either transmission or emission alone.
  • 4.  
    The complete λ = 1–11 μm cool super-Earth emission spectrum had insufficient signal-to-noise for retrievals. There simply is not enough flux contrast Fp/F* for useful emission spectra from this system when a single secondary eclipse is observed at each wavelength. Photometric filter observations may be more useful for constraining the planet's properties. Small (R ≲ 2 R), cool (T < 700 K) planets will need host stars with K ≲ 8.5 mag and/or spectral types later than M0 V for useful emission spectra of single secondary eclipses.
  • 5.  
    The molecular mixing ratios retrieved from λ = 1–5+ μm JWST transmission spectra will provide derived [Fe/H] values with an uncertainty of 0.5 dex (a factor of 3) or better for the clear atmospheres of the hot and warm planets studied. This is adequate for evaluating systematic differences in metallicity with planet mass and determining whether this function is substantively similar to or different from our solar system.
  • 6.  
    Carbon-to-oxygen ratios derived from the retrieved molecular mixing ratios are constrained to better than 0.2 dex (a factor of 1.6) for the hot Jupiter system when using λ ≥ 5 μm transmission or emission spectra. H2O mixing ratio values and retrieved atmospheric temperatures can also be used to distinguish between high (1.0) and solar (0.55) C/O values for hot planets (and with only λ ≤ 2.5 μm transmission data in many cases), sufficient for assessing whether they formed interior to or exterior of the H2O ice lines in their protoplanetary disks. Transmission spectra constrain C/O better than emission spectra for the warm Neptune and sub-Neptune systems. λ ≥ 5 μm spectra provide the best constraints for these warm planets, but λ = 1–2.5 μm data give good results for clear solar atmospheres.
  • 7.  
    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 &Yung (2013). However, observing many planets that span a range of effective temperatures should permit us to identify at which temperatures the molecular abundances deviate from equilibrium. The ensemble of retrieved mixing ratio values and uncertainties of the planets studied here can be used to assess how well each species can be detected in observations of real planets that may have different compositions (e.g., the sensitivity to CO or CO2 in cool planets). The detection limits presented here should be scaled to the S/N of any real data, and similar wavelength ranges should be used in such comparisons.
  • 8.  
    We do not know at this time how well S/N and retrieval uncertainties will improve with the binning or co-addition of data beyond the single-transit or eclipse observations simulated here. We hope and expect that this work will be useful for planning early JWST observations, but actual on-orbit performance must be measured to make better predictions.

We are grateful to L. Albert, J. Barstow, J. Bean, S. Birkmann, J. Bouwman, R. Doyon, P. Ferruit, Th. Henning, L. Kreidberg, P.-O. Lagage, N. Lewis, M. Marley, and the JWST NIRCam and MIRI instrument teams for helpful science discussions, feedback, and information on instrument performance. We also thank the referee I. Crossfield whose numerous insightful recommendations and comments allowed us to substantially improve the paper. This research has made use of the Exoplanet Orbit Database and the Exoplanet Data Explorer at exoplanets.org. The simulations for this research were carried out on the UCSC supercomputer Hyades, which is supported by National Science Foundation (award number AST-1229745) and University of California, Santa Cruz. T.P.G. acknowledges support from the NASA JWST Project and Program for this work via WBSs 411672.04.01.02 and 411672.05.05.02.02. M.R.L. acknowledges support provided by NASA through Hubble Fellowship grant #51362 awarded by the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., for NASA, under the contract NAS 5-26555. J.J.F. acknowledges the support of NSF grant AST-1312545.

Facility: JWST - James Webb Space Telescope

Footnotes

  • We note that thermochemical equilibrium generally does not produce mixing ratio profiles that are constant with altitude. However, for the dominant carbon-bearing species and H2O, constant with altitude generally does occur in equilibrium. For the less dominant species we choose a representative value along a non-uniform vertical profile.

  • 10 

    It is impossible to generate a pure H2O atmosphere thermochemically by scaling the solar elemental abundances. The metallicity value here is determined by taking 0.5 (the ratio of metals (O) to hydrogen (2H)) and dividing that by the value of solar metal fraction of 8.5 × 10−4.

  • 11 

    Instead of using the Differential Evolution Markov chain Monte Carlo (MCMC) as in Line et al. (2013a, 2014b), we use the EMCEE routine of Foreman-Mackey et al. (2014).

  • 12 

    We have also experimented with the centered-log transform (Benneke & Seager 2012) for both solar and HMMW atmospheres and found no significant qualitative differences.

  • 13 
  • 14 

    Also, the true metallicities for the emission spectra are slightly lower than for transmission as we did not include N2 in the total metallicity calculation.

Please wait… references are loading.
10.3847/0004-637X/817/1/17