A publishing partnership

The Detectability and Constraints of Biosignature Gases in the Near- and Mid-infrared from Transit Transmission Spectroscopy

, , , , , , and

Published 2020 February 21 � 2020. The American Astronomical Society. All rights reserved.
, , Citation L. Tremblay et al 2020 AJ 159 117 DOI 10.3847/1538-3881/ab64dd

Download Article PDF
DownloadArticle ePub

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

1538-3881/159/3/117

Abstract

The James Webb Space Telescope is expected to revolutionize our understanding of Jovian worlds over the coming decade. However, as we push toward characterizing cooler, smaller, terrestrial-like planets, dedicated next-generation facilities will be required to tease out the small spectral signatures indicative of biological activity. Here, we evaluate the feasibility of determining atmospheric properties, from near-to-mid-infrared transmission spectra, of transiting temperate terrestrial M-dwarf companions. Specifically, we utilize atmospheric retrievals to explore the trade space between spectral resolution, wavelength coverage, and signal-to-noise on our ability to both detect molecular species and constrain their abundances. We find that increasing spectral resolution beyond R = 100 for near-infrared wavelengths, shorter than 5 μm, proves to reduce the degeneracy between spectral features of different molecules and thus greatly benefits the abundance constraints. However, this benefit is greatly diminished beyond 5 μm as any overlap between broad features in the mid-infrared does not deconvolve with higher resolutions. Additionally, our findings revealed that the inclusion of features beyond 11 μm did not meaningfully improve the detection significance or the abundance constraints results. We conclude that an instrument with continuous wavelength coverage from ∼2 to 11 μm, spectral resolution of R ≃ 50–300, and a 25 m2 collecting area, would be capable of detecting H2O, CO2, CH4, O3, and N2O in the atmosphere of an Earth-analog transiting a M dwarf (magK = 8.0) within 50 transits, and obtain better than an order-of-magnitude constraint on each of their abundances.

Export citation and abstract BibTeX RIS

1. Introduction

Characterizing the climate and composition of terrestrial atmospheres is a primary goal of exoplanet science over the next decades (NAS 2020 Vision6 ). Temperate terrestrial exoplanets are, of course, of great interest due to their possible astrobiological implications, namely, their potential to develop and maintain volatile-rich secondary atmospheres capable of supporting biology. As Earth is the only known planet to host life, it is straightforward that we would begin our search for extraterrestrial life by considering the conditions on Earth as the primary conditions that must exist for the development of life as we know it.

The principal gases that comprise Earth's atmosphere (nitrogen/nitrous oxide, water, carbon dioxide, oxygen/ozone, and methane) are generally regarded as bio-indicators, meaning that their presence is indicative of potentially habitable atmospheric conditions. Biosignatures are those observables that are representative of biological processes affecting the planet's atmospheric chemistry (i.e., the presence of life). Catling & Kasting (2007) showed that all of the bulk gases in Earth's atmosphere, excluding inert gases, are influenced by biogenic processes. Life on Earth is the primary contributor to the global chemical disequilibrium (Krissansen-Totton et al. 2016), and for this reason disequilibrium is presumed to be a robust biosignature (Sagan et al. 1993; Léger 2000; Cockell et al. 2009; Kasting 2009; Seager & Deming 2010; Seager 2014; Seager & Bains 2015). The combinations of molecules responsible for Earth's disequilibrium are O3+CH4 and O3+N2O (Krissansen-Totton et al. 2016). If detected at certain abundances, the presence of these molecules simultaneously cannot be plausibly explained by abiotic processes (Léger et al. 1993; Domagal-Goldman et al. 2014; Tian et al. 2014; Harman et al. 2015, 2018; Schwieterman et al. 2016; Meadows 2017; Meadows et al. 2018). As such, a key goal of exoplanet exploration is to quantify the presence of these gases on nearby terrestrial worlds.

Due to the favorable planet-to-star radius ratio and occurrence rates, M-dwarf systems offer the best near-term opportunity for characterizing temperate terrestrial worlds and the possible presence of biosignatures. Ground-based radial velocity surveys and the Transiting Exoplanet Survey Satellite (TESS) are expected to detect dozens of temperate terrestrial-sized (0.8–1.6 R) exoplanets around M dwarf's (Barclay et al. 2018; Kempton et al. 2018; Quirrenbach et al. 2018; Zechmeister et al. 2019), potentially amenable to follow-up with the James Webb Space Telescope (JWST). Numerous previous works have demonstrated that JWST will likely provide astounding constraints on the atmospheric properties of Jovian-to-super-Earth planets (Deming et al. 2009; Benneke & Seager 2012; Barstow et al. 2014; Beichman et al. 2014; Greene et al. 2016; Rocchetto et al. 2016; Batalha et al. 2018). However, it remains an open question as to how well JWST will perform as it is pushed to observe temperate terrestrial bodies with smaller spectral features. Possible limitations may stem from a combination of large detector noise floors (similar detector technology as the Hubble Space Telescope WFC3 and Spitzer IRAC; Greene et al. 2016), saturation limits for the brightest targets (Batalha et al. 2018), and a lack of continuous near-to-mid-infrared wavelength coverage in a single mode (greatly increasing the number of required transits).

The discovery of the TRAPPIST-1 system (Gillon et al. 2016) has lead to a number of works investigating the climate, composition, and the internal structure of temperate M-dwarf worlds (Meadows 2017; Unterborn & Panero 2017; Suissa & Kipping 2018; Turbet et al. 2018; Wunderlich et al. 2019). Complementary to these efforts, recent studies have explored the capabilities of JWST to characterize various atmospheric compositions on TRAPPIST-1 planets through both transmission and emission spectroscopy (Barstow & Irwin 2016; Morley et al. 2017; Batalha et al. 2018; Krissansen-Totton et al. 2018; Lustig-Yaeger et al. 2019; Wunderlich et al. 2019).

Morley et al. (2017) concluded that JWST's NIRSpec G235M/F170LP mode (1.66–3.07 μm) would be capable of rejecting a flat-line spectrum at 5σ confidence within 20 transits for a range of atmospheric compositions. While a flat-line rejection test is valuable, it does not guarantee the ability to detect the presence of any specific molecule with high statistical confidence. Lustig-Yaeger et al. (2019) concluded that transmission spectroscopy with NIRSpec Prism is optimal for detecting the presence of high mean molecular weight atmospheres within ∼12 transits. Additionally, they find that CO2, if present, will be easily detectable regardless of the atmospheric composition and Mid-Infrared Instrument-Low Resolution Spectrometer (MIRI-LRS) may be capable of detecting the 9.6 μm O3 on planet 1e (with a signal-to-noise ratio (S/N) = 3) with greater than 100 transits. Wunderlich et al. (2019) produced self-consistent forward models for the atmosphere of an Earth-like planet around early-to-late M dwarfs. Their "Earth around TRAPPIST-1" model notably predicts a greatly increased temperature in the mid-to-upper atmosphere, inflating the scale height and increasing the detectability of molecular species for transmission spectroscopy with JWST. Furthermore, the predicted chemical evolution indicated a significant enhancement of CH4 and H2O compared to modern Earth and subsequently, that their corresponding spectral features, along with CO2, would be detectable within ∼10 transits when using the appropriate NIRSpec high-resolution filters.

Barstow & Irwin (2016) performed a retrieval analysis on synthetic spectra of the innermost TRAPPIST-1 companions (b, c, d), each with an Earth-like composition atmosphere. Utilizing a combination of NIRSpec Prism and MIRI-LRS, they conclude that O3, at Earth-like abundances, would be detectable on planets 1c and 1d with 30 transits on each instrument. Also utilizing an atmospheric retrieval analysis, Krissansen-Totton et al. (2018) analyzed the detectability of the CO2+CH4 disequilibrium biosignature, suspected to have been present in Earth's early atmosphere when the abundances of each were significantly enhanced compared to modern-day values. They concluded that JWST would be capable of obtaining order-of-magnitude constraints on the mixing ratios of both CH4 and CO2 in an Archean Earth-like atmosphere, in ∼10 transits. Alternatively, the abundance of O3 in a modern Earth-like atmosphere could not be constrained by NIRSpec PRISIM and would be completely unbounded by MIRI-LRS.

While it may be possible for JWST to detect and constrain the dominant molecular species in certain terrestrial atmospheres with advantageous compositions, existing literature does not substantiate its capabilities to detect or constrain the abundances of the five primary bio-indicators in a modern Earth-like atmosphere on an M-dwarf companion.

The primary aim of this work is to develop a baseline understanding of how key observational parameters (spectral resolution, wavelength coverage, and signal-to-noise) of near-to-mid-infrared transmission spectra influence the ability to detect and constrain biologically relevant molecular species in the atmospheres of transiting Earth-like planets orbiting M dwarfs. We leverage powerful Bayesian retrieval tools to obtain both parameter constraints and Bayesian molecular detection significances as a function of the instrumental trades. In Section 2 we describe our model/retrieval and instrumental-trade space setup, Section 3 summarizes the key results, followed by a discussion and key points in Sections 4 and 5.

2. Simulation Setup

Here, we outline the development of our forward model used to compute the synthetic transmission spectra, elaborate on the instrumental-trade space, discuss the details of our instrument noise model, and review our retrieval approach.

2.1. Transmission-forward Model

We simulate the transmission spectrum of a terrestrial planet, with mass and radius equivalent to TRAPPIST-1e, residing within the habitable zone of TRAPPIST-1 with an approximate modern Earth-like atmosphere (observational system setup/signal-to-noise discussed below). We acknowledge, up front, that this may not be a physically plausible scenario due to the vastly different incident spectral energy distribution. However, given the overwhelming number of unknowns involved in self-consistent planetary atmosphere modeling (e.g., star–planet interactions, surface fluxes, formation conditions, bulk elemental composition, 3D atmospheric dynamical effects, cloud microphysics, dynamical evolution/history) we simply choose to treat our atmosphere as Earth-like (as has also been done in previous works; Morley et al. 2017; Krissansen-Totton et al. 2018).

To produce model transmission spectra, we leverage a variant7 of the CHIMERA transmission spectrum routine (Line et al. 2013; Greene et al. 2016; Line & Parmentier 2016; Batalha & Line 2017; Line et al. 2013; Greene et al. 2016; Line & Parmentier 2016; Batalha & Line 2017). Specifically, we re-parameterize the code (Table 1) to make it more amenable for temperate worlds by including as free parameters the constant-with-altitude mixing ratios of H2O, CO2, CO, CH4, O3, N2O and an unknown background gas with a free-parameter molecular weight (taken to be Earth's N2+O2 value), an isothermal scale-height temperature, planetary radius at the surface (or scaling thereof), and an opaque gray cloud-top pressure (nominal truth values are given in Table 1). We mimic refraction with this cloud-top pressure (here, 0.56 bars) using the prescription from Robinson et al. (2017), but otherwise assume a cloud-free atmosphere (the refractive layer is in effect, mimicking a cloud, as well as its inclusion as a free parameter). CHIMERA uses correlated-K opacities (computed at the given constant resolving powers below), here derived from a grid of pre-computed line-by-line (≤0.01 cm−1, 70–410 K, 10−7–30 bar) cross-sections generated with the high-resolution transmission molecular absorption database (HITRAN) Application Programming Interface (HAPI) routine (Kochanov et al. 2016) and the HITRAN2016 line database (Gordon et al. 2017). We do not include collision-induced opacities, though they may spectrally present themselves throughout the mid-infrared (Schwieterman et al. 2015).

Table 1.  The 10 Free Parameters, the Model Values, and Their Associated Uniform Priors Used in our CHIMERA Radiative Transfer and Retrieval Code

  Model Value Uniform Prior
log(H2O) −5.5 [−12, 0]
log(CO2) −3.45 [−12, 0]
log(CH4) −6.3 [−12, 0]
log(O3) −6.5 [−12, 0]
log(N2O) −6.3 [−12, 0]
log(CO) −7.0 [−12, 0]
Tiso 280 K [100, 800]
xRP 0.918 R [0.5, 1.5]
log(CTP) −0.25 [−6, 2]
Bkg MMW 28.6 [2.0, 44.0]

Download table as:  ASCIITypeset image

We benchmarked our transmission-forward model against those available from the Virtual Planetary Laboratory8 (Figure 1).

Figure 1.

Figure 1. Validation of our transmission-forward model (at R = 100) against the Virtual Planetary Laboratory (VPL) Earth model3 utilizing the same temperature pressure and gas mixing ratio profiles. The spectra were scaled to a 1 R around a 0.15R star. For reference, overlaid are 5 ppm error bars.

Standard image High-resolution image

2.2. Parameter Estimation and Model Selection Approach

We perform Bayesian parameter estimation and model selection, on the simulated data sets described below, using the PyMultiNest routine (Buchner et al. 2014) following the methods described in Benneke & Seager (2013). We initialized our retrievals with 3000 live points. An advantage of nested sampling algorithms is the ease of evidence computation, which can be used to assess model complexity. We use Bayesian nested model comparison (by removing one gas at a time) to determine the detection significance of each molecular species (Trotta 2008; Benneke & Seager 2013), and it is what we utilize as a metric for assessing instrument performance. An advantage of utilizing the detection significance via the Bayesian evidence over more traditional methods (e.g., line or band height above a continuum relative to the noise) is that it fully utilizes all of the spectral information included in all of the relevant bands as well as encompasses the model degeneracies.

2.3. Simulated Spectrograph and Radiometric Noise Model

We use the JWST radiometric noise model described in Greene et al. (2016) to produce spectrophotometric uncertainties. The collecting area of 25 m2 was retained to reflect the identical collecting area of JWST in order to provide a baseline with which to easily scale the results of this study. Modifications were made to parameters of the code to reflect the expected performance of a next-generation mid-infrared telescope. These parameters included a 2.85–30.0 micron wavelength range (HgCdTe detectors below 10.5 microns and SiAs detectors above 10.5 microns), a zodiacal background estimated by Glass & Blair (2015) and Glass et al. (2015), and an intrinsic resolving power of 300. These modifications reflect the capabilities of new detector technologies discussed by Matsuo et al. (2018).

Our input stellar spectrum derives from the PHOENIX stellar models using the PySynPhot software package (STScI Development Team 2013). We adopted the values for TRAPPIST-1 (T = 2550K,M/H = 0.40, log(g) = 4.0) and a stellar radius of 81373.5 km from Gillon et al. (2016). Additionally, we have scaled the stellar spectrum to a K-band magnitude of 8.0 rather than magK = 10.3, TRAPPIST-1's native magnitude. While this may be representative of an optimistic scenario, magK = 8.0 is the mean magnitude of the brightest 10 M-dwarf stars in the Barclay catalog of the predicted TESS yield (Barclay et al. 2018).

2.4. Observational Parameter Space

We produce synthetic transmission spectra of our Earth-like atmosphere model over a grid of spectral resolution and bandpass choices. The nominal wavelength coverage was chosen to cover the near- and mid-infrared regimes spanning 1–30 μm, shown in Figure 2. We divide this wavelength range into eight separate regimes to explore the influence of additional bands of a given molecule and their overlap with features of other molecules. Additionally, we test low and moderate spectral resolutions ranging from R = 30 to R = 300. The spectral resolution (λ/δλ) dictates the degree to which individual molecular bands can be resolved and overlap degeneracies broken. Figure 3 compares the spectra at the resolutions explored in this work. We simulate different resolutions within each spectral band in order to explore the resolution trade along with each wavelength regime. This produces a grid of 23 test cases (shown in Table 2) to evaluate through our retrieval algorithm described in 2.2.

Figure 2.

Figure 2. Wavelength region explored in this work (here, shown at R = 50). The colored bands indicate the presence, and approximate width, of a spectral feature for a given molecule. We explore constraints obtainable over several wavelength bands within this region (in Table 2).

Standard image High-resolution image
Figure 3.

Figure 3. Influence of resolution on spectral features. Lower resolution results in the blending of features, causing a loss of information. The vertical lines at 3, 5, and 11 μm denote the division of the wavelength regimes explored in this study.

Standard image High-resolution image

Table 2.  A Breakdown of the 23 Test Cases Evaluated in this Study as a Combination of Wavelength Coverage and Spectral Resolution

Short Wavelength Long Wavelength Spectral Resolution
Boundary Boundary  
1 μm 5 μm 300
  100
  50
  11 μm 100
  50
  30 μm 100
  50
  30
3 μm 5 μm 300
  100
  50
  11 μm 300
  100
  50
  30 μm 100
  50
  30
5 μm 11 μm 100
  50
  30
  30 μm 100
  50
  30

Download table as:  ASCIITypeset image

In addition to resolution and wavelength coverage, we also explored the signal-to-noise trade, parameterized here as the number of transits (1–100), which can be considered as a proxy for the mirror diameter or source magnitude, where the noise on a single transit is defined by the nominal noise model setup described in Section 2.3. Utilizing this trade, we explored the minimum number of transits necessary to achieve a threshold 3.6σ confidence level detection of each molecule. We provide Equations (1) and (2) as a means of converting our retrieval results to stars of differing magnitudes or for different sized collecting mirrors.

Equation (1)

where Teff/Tnom is the ratio determining how many more or less transits are equivalent to a change from the native magnitude (Mnom = 8.0) to a new magnitude (Meff).

Equation (2)

where Teff/Tnom is the ratio of transits equivalent to a change from the native collecting area (Anom = 25 m2) to a new mirror size (Aeff).

3. Retrieval Results

For the majority of our test cases, the inclusion of a broader bandpass decreases both the necessary observation time to claim a statistically significant detection and reduces the uncertainty on the abundance constraints. Therefore, with few molecule-specific exceptions, we find that altering the choice of wavelength coverage has a more prominent benefit for both our detection significance values and abundance constraints than the choice of spectral resolution (Tables 3 and 4).

Table 3.  Number of Transits Required to Detect Each Molecule to 3.6σ Confidence for Each of Our Test Cases

Bandpass Resolution H2O CO2 CH4 O3 N2O
1–5 μm 300 18.1 <1 38.5 7.7
  100 23.3 <1 54.4 9.2
  50 39.2 <1 14.3
1–11 μm 100 6.5 <1 44 8.5 5.7
  50 7.3 <1 52.1 9.4 7.4
1–30 μm 100 6.3 <1 43.8 8.3 5.5
  50 7.0 <1 50 9.1 7.2
  30 8.5 1.13 10.9 9.0
3–5 μm 300 1.03 54.9 11.2
  100 1.08 14.4
  50 1.13 23.8
3–11 μm 300 11.9 <1 34.3 8.9 5.6
  100 15.4 1.07 61.7 10.5 7.9
  50 19.6 1.11 12.9 10.6
3–30 μm 100 14.4 <1 60 10.2 7.7
  50 17.4 <1 11.8 9.6
  30 21.1 1.15 14.7 12.2
5–11 μm 100 33.1 14.9 94.0
  50 50 18.2
  30 67.2 24.3
5–30 μm 100 22.5 6.75 12.9 72.8
  50 29.6 6.88 15.1
  30 36.8 6.96 18.2

Note. A ... indicates that achieving a confidence level of 3.6σ was not possible within 100 transits.

Download table as:  ASCIITypeset image

Our detection significance analysis directly compares each bandpass and resolution combination by determining the number of transits required to achieve a 3.6σ detection for a specific molecule. Table 3 displays these required number of transits for each molecule over each combination of resolution and wavelength coverage.

Table 4 and Figure 4 summarize the constraints for each molecule within each of our 23 test cases at 50 transits. Figures 59 within each of the following sections reveal the influence that an increasing number of transits has on the abundance constraints, for each molecule individually at R = 100. The following subsections provide an in-depth analysis of each molecule and the most relevant transitions that influence our abundance constraints and detection significance results.

Figure 4.

Figure 4. Abundance constraints for each molecule within three key bandpasses (1–5 μm, 3–11 μm, and 5–30 μm at R = 100) as a function of the number of observed transits. The height of each histogram window has been fixed to a constant value to allow for a direct comparison of the shape of the marginalized posterior probability distributions for each gas. The "hard edge" of some distributions, near low abundances, indicates those that extend beyond the prior range.

Standard image High-resolution image
Figure 5.

Figure 5.�Compares the improvements of the abundance constraints for H2O, with increasing numbers of transits at three resolutions for the two key bandpass choices.

Standard image High-resolution image
Figure 6.

Figure 6.�Compares the improvements of the abundance constraints for CO2, with increasing numbers of transits at three resolutions for the two key bandpass choices.

Standard image High-resolution image
Figure 7.

Figure 7.�Compares the improvements of the abundance constraints for CH4, with increasing numbers of transits at three resolutions for the two key bandpass choices.

Standard image High-resolution image
Figure 8.

Figure 8.�Compares the improvements of the abundance constraints for O3, with increasing numbers of transits at three resolutions for the two key bandpass choices.

Standard image High-resolution image
Figure 9.

Figure 9. Compares the improvements of the abundance constraints for N2O, with increasing numbers of transits at three resolutions for the two key bandpass choices.

Standard image High-resolution image
Figure 10.

Figure 10. Relationship between the detection significance and abundance constraint. The numbers on each data point represent the number of transits observed to achieve those respective results. The number of transits for the N2O (orange) are identical to those of H2O (blue), conveniently adjacent here. These results are representative of an instrument with 1–30 μm coverage, R = 100 for a stellar mK = 8, and a mirror aperture of 25 m2. Unsurprisingly, an increase in detection significance is correlated with higher precision. However, the mapping between detection significance and constraint is molecule specific.

Standard image High-resolution image

Table 4.  Logarithmic Constraints on the Vertical Mixing Ratios for Each Molecule, at 50 Transits, for Each of Our Test Cases

Bandpass Resolution H2O CO2 CH4 O3 N2O
1–5 μm 300 0.77 0.81 0.63 0.65 0.77
  100 0.99 1.01 0.81 $\uparrow $ 0.97
  50 1.19 1.24 1.03 $\uparrow $ 1.19
1–11 μm 100 0.79 0.77 0.65 0.60 0.74
  50 0.99 0.93 0.81 0.75 0.92
1–30 μm 100 0.77 0.74 0.63 0.59 0.72
  50 0.97 0.91 0.80 0.73 0.89
  30 1.15 1.08 0.99 0.87 1.07
3–5 μm 300 $\uparrow $ 1.31 1.04 0.95 1.16
  100 $\uparrow $ 1.44 1.32 $\uparrow $ 1.27
  50 $\uparrow $ 1.86 $\uparrow $ $\uparrow $ 1.65
3–11 μm 300 0.95 0.91 0.79 0.68 0.85
  100 1.15 1.07 0.99 0.82 1.02
  50 1.37 1.29 1.30 0.99 1.23
3–30 μm 100 1.10 0.74 0.94 0.79 0.99
  50 1.33 0.91 1.21 0.96 1.21
  30 1.54 1.08 1.52 1.10 1.38
5–11 μm 100 1.89 $\uparrow $ 1.41 1.70
  50 2.05 1.58 2.01
  30 2.10 1.58 2.03
5–30 μm 100 1.70 1.61 $\uparrow $ 1.23 1.50
  50 1.80 1.68 $\uparrow $ 1.29 1.69
  30 1.92 1.74 $\uparrow $ 1.33 1.78

Note. The values reported are the half-width of the posterior distribution, therefore, 1.0 dex indicates the ability to constrain the abundance to within an order of magnitude higher or lower (e.g., ±1 dex) than the retrieved value. The $\uparrow $ marker denotes an upper limit and the ... represents that the parameter was entirely unconstrained by the retrieval. The posteriors for all of the R = 50 and R = 100 test cases are provided on Zenodo (doi:10.5281/zenodo.3630884).

Download table as:  ASCIITypeset image

3.1. Water (H2O)

Water has spectral features spanning nearly the entire wavelength range in this study, with the most prominent features centered at 2.65 and 6.3 μm. Additionally, there are two smaller features extending into the near-infrared at 1.4 and 1.85 μm as well as a series of features upwards of 20 μm. The most relevant comparison is between the near-infrared (which we define as the 1–5 μm bandpass) and the mid-infrared (5–11 μm). In isolation from other water features in the 5–11 μm region, the broad 6.3 μm feature produces comparatively underwhelming results (for both detection significance and abundance constraints) likely due to the limited contribution from the few other prominent features in this bandpass (Benneke & Seager 2012). In contrast, the 1–5 μm region encompasses the prominent 2.65 μm feature as well as the two smaller features. The precision on the abundance constraints improves by approximately 60% both when increasing the resolution from R = 50 to R = 100 and again from R = 100 to R = 300. However, it is worth noting that these near-infrared features are very narrow and are thus significantly more sensitive to changes in the choice of spectral resolution. At a resolution of R = 100, a single H2O feature in the near-infrared consists of between 10 and 20 resolution elements compared to 44 for the 6.3 μm feature.

While the near-infrared (1–5 μm bandpass) produces the tightest abundance constraints, it does not provide the greatest detection significance values given the same spectral resolution and number of transits (refer to Table 3 for a comparison of the detection significance results). By comparison, the 3–11 μm range (which does not include additional H2O features) requires 33.9% less observation time at R = 100 and 50.0% less at R = 50 than the 1–5 μm bandpass. Alternatively, by looking further into the mid-infrared, the 5–30 μm bandpass offers a 3.4% decrease at R = 100 and a 24.5% decrease at R = 50 compared to the 1–5 μm bandpass.

3.2. Carbon Dioxide (CO2)

Due to its prominent spectral features across most of the near- and mid-infrared, CO2 is the easiest molecule to detect within each of our bandpass choices, with the exception of 5–11 μm. CO2 possess four prominent spectral features centered at 2.0, 2.7, 4.3, and 15um. The 5–11 μm region is the only bandpass choice that does not include a spectral feature for CO2. Any other choice of wavelength coverage and spectral resolution evaluated in our study proved to be sufficient to detect CO2 in less than seven transits. The inclusion of the 4.3 μm feature ensures a 3.6σ detection within the observation time of approximately a single transit.

Despite the strong detection of CO2, the abundance constraints are, ironically, not as precise as would be expected. This is largely due to the fact that deriving narrow abundance constraint relies on a change in the strength of the spectral features. Line & Parmentier (2016) derive the function for spectral modulation with respect to wavelength for multiple absorbing molecular species, reproduced here for convenience (Equation (3)).

Equation (3)

where λ/ is the wavelength-dependent slope of the transmission spectra, H is the scale height given by $\tfrac{{k}_{b}T}{\mu g}$, and ${\sigma }_{\lambda ,i}$ and ξi are the absorption cross-section and abundance, respectively, of a given molecular species.

In the case of CO2, the ξ1σλ,1 term is so much greater than the other absorbers that the spectral modulation due to CO2 becomes insensitive to even large changes in the abundance. The retrieval thus determines that a wide range of abundance values can reproduce the shape of the observed features within the error bars. Therefore, the only means to narrow the constraint is to increase the number of observed transits.

At 50 transits, the 1–5 μm bandpass produces approximately a single order-of-magnitude (1.01 dex) constraint at R = 100 or 0.81 dex at R = 300. The abundance constraints produced by the 1–5 μm bandpass cases at both R = 50 and R = 100 are comparable to those of both the 3–11 μm and 3–30 μm bandpasses.

3.3. Methane (CH4)

Due to its low abundance in the modern Earth atmosphere, CH4 is difficult to detect and ultimately determines the lower limit on the observing time required to detect all five bio-indicator molecules. Despite the larger number of transits required to detect CH4, we can place the best constraints on its abundance. Indeed, it is notable that our results indicate that for the 3–11 μm at R = 100 case, a 3.6σ detection is only possible after 61.7 observed transits while at only 50 transits we can place an order-of-magnitude constraint on its abundance. This appears initially paradoxical, until we examine the underlying dependencies of these two different results. Effectively, the detectability (via the Bayes factor) depends on the amplitude of the spectral features with respect to the continuum or adjacent features of other molecules (e.g., the wavelength-dependent derivative of transit depth shown in Equation (3)). The mixing ratio constraints on the other hand are dictated by the derivative of the transit depth with respect to a given species abundance (or rather, the inverse, e.g., Line et al. 2012), with each species having varying sensitivity. Therefore, one could observe small features resulting in a low detectability (as in the case of CH4) while having a large abundance derivative, resulting in a tight constraint. Figure 10 illustrates the molecule-specific relationship between the abundance constraints (log) and detection significance under the 1–30 μm, R = 100 scenario for varying numbers of transits. Figure 10 illuminates the key differences between CH4 and CO2 in that while CO2 has a high detection significance, it does not necessarily have a precise constraint. Alternatively, even low confidence detections of CH4 can yield precise abundance constraints.

At Earth-like abundances, CH4 has only three discernible spectral features in the near- and mid-infrared: 2.3, 3.3, and 7.6 μm. The most prominent of these is centered at 7.6 μm and directly overlaps with a larger N2O feature centered at 7.7 μm. Bandpass choices that exclude additional methane features will lack the ability to overcome this degeneracy. Fortunately, the 3.3 μm (ν3) feature provides an unambiguous marker for the presence of CH4. We find that the inclusion of this feature is crucial to detecting CH4 as neither the 5–11 μm nor the 5–30 μm bandpasses are capable of detecting methane within 100 transits, despite the presence of the 7.6 μm (ν4) feature. There is an additional feature at 2.3 μm which, when combined with the 3.3 μm band, greatly improves the detection significance.

When observing only the 3–5 μm bandpass, at R = 300, CH4 is detectable with 54.9 transits and provides an abundance constraint of 1.04 dex at 50 transits. Maintaining this narrow wavelength coverage and reducing to R = 100, increases the necessary observation time to beyond 100 transits. However, expanding our bandpass either to the near- or mid-infrared regimes proves to provide significant benefits. The near-infrared bandpass (1–5 μm) requires 38.5 transits and provides a 0.63 dex (at 50 transits) abundance constraint or 54.4 transits and 0.81 dex for R = 300 and R = 100, respectively. Alternatively, observing the 3–11 μm bandpass requires 34.3 transits and provides a 0.79 dex (at 50 transits) abundance constraint or 61.7 transits and 0.99 dex for R = 300 and R = 100, respectively.

Based on these results, we find that including the ν3 feature, within the chosen bandpass, is essential to detecting and constraining CH4. In the scenario that spectral resolution is limited to R = 100, the addition of the 2.3 μm feature is incredibly beneficial, providing a 0.5 dex improvement in the abundance constraint and requiring only a third of the observational time. However, while it is certainly ideal to include both features if possible, it is quite noteworthy that the benefit of a higher resolution (R = 300) choice, in the absence of the 2.3 μm feature, outweighs the benefit of including this feature at a lower resolution (R = 100).

3.4. Ozone (O3)

Ozone is different from the other molecular species we have addressed thus far, in that its most prominent features extend further into the mid-infrared. The three primary features for O3 are at 4.75, 9.6, and 14.2 μm with the 9.6 μm feature being the strongest. Conveniently, due to the location of these spectral features and our bandpass choices, we can evaluate the benefit of each additional feature, individually, on the retrieval results.

Isolating the 9.6 μm feature, in the 5–11 μm bandpass, requires a relatively low number of transits at only 14.9 and 18.2 transits for resolutions of R = 100 and R = 50, respectively. However, the abundance constraints for this bandpass exceed an order of magnitude at 1.41 dex (R = 100) and 1.58 dex (R = 50). If we choose to observe further into the mid-infrared (5–30 μm), including the features at 9.6 and 14.7 μm, we notice only slight reductions in the required observation time and maintain larger than an order-of-magnitude constraint on both abundance values. The one noticeable benefit of including these longer wavelengths is that less than 20 transits are required to detect O3 even at a resolution of R = 30. Alternatively, extending the bandpass to 3–11 μm to include the 4.75 μm feature results in a reduction of the necessary observation time to 10.5 transits for a resolution of R = 100 (12.9 for R = 50) and improves our abundance constraints by nearly a factor of four (0.82 and 0.99 dex) for both R = 100 and R = 50. It is worth noting that the 4.75 μm feature has significant overlap with a minor CO2 feature, thereby causing a degeneracy if observed without additional O3 features. We conclude that the optimal wavelength coverage for detecting and constraining O3 is the 3–11 μm bandpass choice. Additionally, this bandpass would only require a resolution of R = 50 to place an order-of-magnitude constraint on the volume mixing ratio for O3.

3.5. Nitrous Oxide (N2O)

Similar to O3, the most prominent features for N2O extend out into the longer wavelengths. N2O's strongest spectral features are located at 4.5 and 7.7 μm, each with an adjacent smaller feature at 3.9 μm and 8.6 μm, respectively. Although there is a broad feature at 17 μm, it is overlapped by a much larger CO2 feature.

Due to the large degree of degeneracy with the 7.6 μm CH4 feature, isolating the 7.7 and 8.6 μm features in the 5–11 μm bandpass proves to be inadequate. At a spectral resolution of R = 100, the 5–11 μm bandpass requires nearly 12 times as many transits (∼94 transits) as the 3–11 μm bandpass would require to obtain a 3.6σ detection. Extending to 30 μm incorporates the 17 μm feature and only reduces the observational time to 72.8 transits while the constraints still exceed 1.50 dex.

We find that the 3–5 μm bandpass is sufficient to detect N2O at 14.4 transits (R = 100) or 23.8 transits (R = 50), making it one of the only two molecular species that can be constrained in that bandpass within relatively few transits. However, that region is too narrow to provide adequate abundance constraints, resulting in a 1.27 dex constraint at 50 transits. If we expand to either the 1–5 μm or the 3–11 μm bandpass, our required number of transits drops to 9.2 and 7.9, respectively. The corresponding constraints at 50 transits are 0.97 dex (1–5 μm) and 1.02 dex (3–11 μm), roughly an order of magnitude for both. The 3–11 μm region would thus allow for a 3.6σ detection in fewer transits than would be required for either O3 or H2O within the same regime. Likewise, the 1–5 μm bandpass requires only 39.5% of the observational time, to detect N2O, as needed to detect H2O in the same regime. With comparable abundance constraints, the 3–11 μm bandpass proves to be the optimal bandpass, requiring only 74%–86% (R = 50 to R = 100) of the observation time as the 1–5 μm bandpass choice.

4. Discussion

Unsurprisingly, we find that obtaining the best detection significance values and abundance constraints for all of the molecules of interest are achieved by including the broadest possible wavelength coverage at the highest possible spectral resolution. Our primary goal, however, was to determine how the detection significance and abundance constraints behave as a function of the continuous wavelength coverage and spectral resolution. In the previous section, we elaborate on the effect that additional wavelength coverage and finer resolution have on the retrieval results for each molecule individually. Below we summarize our key findings.

  • 1.  
    When considering the contribution of additional wavelength coverage to the capability of an instrument with a continuous infrared bandpass, the most crucial spectral features (for H2O, CO2, CH4, O3, N2O, and CO) in transmission spectra do not extend beyond 11 μm.
  • 2.  
    Utilizing a spectral resolution greater than R = 100 proves to be significantly beneficial for resolving features at wavelengths shorter than 5 μm. Redder wavelengths offer notably broader features, easily resolved with a resolution of R = 50.
  • 3.  
    Degeneracies caused by significantly overlapping spectral features are most easily resolved by including an additional feature for one or both molecules.

A brief summary of the crucial molecule-specific takeaways is as follows.

  • 1.  
    When attempting to detect and constrain CH4, the observation of the 3.3 μm feature is essential. If one must opt for lower resolutions (R = 100), including the 2.3 μm feature is incredibly valuable. While it is certainly ideal to include both features, it is crucial to acknowledge that the benefit of a higher resolution (R = 300), in the absence of the 2.3 μm feature, outweighs the benefit of including this feature at a lower resolution (R = 100).
  • 2.  
    The results of this study validate an alternative to the observing strategies that are currently available for H2O, which are limited to the weak near-infrared features. We find that the more prominent unobscured spectral feature centered at 6.3 μm, combined with the dense cluster of features from other molecules at 3–5 μm, provides a significant decrease in the necessary observing time to detect H2O compared to the 1–5 μm bandpass.
  • 3.  
    CO2, due to its very prominent features, is easily detectable even only observing the 4.3 μm feature. However, increasing the wavelength coverage has statistically significant effects on the abundance constraints regardless of whether the broader bandpass includes additional CO2 features.
  • 4.  
    Given a mK = 8.0 source star, we find that utilizing an instrument with wavelength coverage from 3 to 11 μm, a spectral resolution of only R = 50, and a 25 m2 collecting area is sufficient to detect ozone abundances representative of modern-day Earth with less than 13 transits and to constrain the abundance value within an order of magnitude at 50 transits.
  • 5.  
    All of the N2O features from 1 to 30 μm are partially or completely degenerate with features from CO2 or CH4. However, under the detector setup evaluated in this study, observing only the features in the 3–5 μm regime allows for a statistically significant detection in less than 15 transits. The 1–5 μm and the 3–11 μm bandpasses offer comparable improvements to both the detection significance and abundance constraints.

The ability to detect and constrain the five biologically relevant molecular species in the atmosphere of an Earth-analog should be the benchmark by which the next-generation of exoplanet observatories are designed. The detection of significant amounts of CH4 in tandem with O3 and/or N2O would constitute a promising indicator of biogenic origins. However, as it has been laboriously stated, CH4 is the most challenging molecule to detect in this study and will prove to be the limiting variable in the development of a detector for these purposes. Unlike the other molecules considered here, CH4 does not possess broad unobscured features in the mid-infrared and therefore does not have the advantage of relaxed spectral resolution requirements. It is important to acknowledge, however that this is not an argument against observing mid-infrared wavelengths, which have been shown to be immensely valuable in the detection and constraint of all of the other molecular species in this study. Rather, our intention is to highlight that the choice of wavelength coverage for a next-generation near-to-mid-infrared spectrometer, particularly on the blue end of the bandpass, must be influenced by one or both of the near-infrared CH4 features. To that end, we conclude that a near-to-mid-infrared spectrometer, on board a next-generation space-based telescope (with a collecting area of 25 m2), boasting a bandpass including features between ∼2 and 11 μm and a resolution of R ≃ 50–300 would prove capable of detecting and constraining all of the key bio-indicator gasses in Earth's atmosphere. While the 2–11 μm bandpass was not explicitly probed in this study, our retrieval results indicate that it would provide 3.6σ detections in less transits and narrower abundance constraints than either the 1–5 μm or 3–11 μm bandpasses studies here.

5. Conclusion

Our work has been motivated by the desire within the exoplanet community to characterize the atmospheres of temperate terrestrial planets beyond our solar system. The detection of molecular biosignatures on these small rocky planets looms just beyond the capabilities of current space-based telescopes. A future facility with broad continuous wavelength coverage in the near- and mid-infrared combined with detectors capable of driving observational noise down to the astrophysical noise floor will have what it takes to comprehensively probe the climate and composition of terrestrial exoplanet atmospheres.

The 2020 decadal survey will consider four flagship mission concepts including one of particular interest to this study, the Origins Space Telescope. The Origins concept includes a near-to-mid-infrared spectrograph boasting an impressive 2.8–20 μm continuous bandpass ranging from R ≃ 50–300 with detectors designed to reduce instrument noise to approximately 5 ppm (Origins Space Telescope Mission Concept Study Report9 ). Although it excludes the 2.3 μm feature for CH4, the design of this instrument strongly aligns with the findings in this work for the optimal observational setup for both detecting the presence and constraining the abundances of Earth-like biosignatures.

We look to expand upon this work by testing a subset of this analysis to a broader range of atmospheric compositions in order to determine the possibility of distinguishing several plausible atmospheric compositions on known terrestrial planets and those soon to be discovered by TESS. Additionally, we acknowledge the benefit of thermal emission spectroscopy for characterizing the thermal structure of exoplanetary atmospheres and providing additional information on the molecular abundances. Therefore, we aim to explore the effect of these observational parameters on a similar grid of synthetic thermal emission spectra. To this end, we anticipate that the synthesis of transmission and thermal emission as well as reflected light spectroscopy to be essential to the comprehensive understanding of these exoplanetary environments.

We would like to thank Tom Greene for providing us with his noise model. M.R.L. and L.T. acknowledge the NASA XRP award NNX17AB56G for supporting this work. This work benefited from the 2018 Exoplanet Summer Program in the Other Worlds Laboratory (OWL) at the University of California, Santa Cruz, a program funded by the Heising-Simons Foundation. We would also like to acknowledge Origins Space Telescope Exoplanets Working Group for supporting travel related to this work.

Part of the research was carried out at the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration.

Software: PySynPhot (STScI Development Team 2013), CHIMERA (Line et al. 2013; Greene et al. 2016; Line & Parmentier 2016; Batalha & Line 2017), PyMultiNest (Buchner et al. 2014).

Appendix

Figure 11 summarizes via a corner plot the constraints and degeneracies among all of the retrieved model parameters for a representative case (3–11 μm, R = 100, 50 transits). Corner plots for all of the R = 50 and R = 100 scenarios are provided on Zenodo (doi:10.5281/zenodo.3630884).

Figure 11.

Figure 11. Corner plot for the 3–11 μm at R = 100 test case. The posterior probability distributions for all of the parameters in our model are shown. All of the corner plots for the R = 50 and R = 100 scenarios are provided on Zenodo (doi:10.5281/zenodo.3630884).

Standard image High-resolution image

Footnotes

Please wait… references are loading.
10.3847/1538-3881/ab64dd