GLOBAL MAPPING OF EARTH-LIKE EXOPLANETS FROM SCATTERED LIGHT CURVES

and

Published 2010 August 19 © 2010. The American Astronomical Society. All rights reserved.
, , Citation Hajime Kawahara and Yuka Fujii 2010 ApJ 720 1333 DOI 10.1088/0004-637X/720/2/1333

0004-637X/720/2/1333

ABSTRACT

Scattered lights from terrestrial exoplanets provide valuable information about their planetary surface. Applying the surface reconstruction method proposed by Fujii et al. to both diurnal and annual variations of scattered light, we develop a reconstruction method of land distribution with both longitudinal and latitudinal resolutions. We find that one can recover a global map of an idealized Earth-like planet on the following assumptions: (1) cloudlessness, (2) a face-on circular orbit, (3) known surface types and their reflectance spectra, (4) lack of atmospheric absorption, (5) known rotation rate, (6) a static map, and (7) the absence of a moon. Using the dependence of light curves on planetary obliquity, we also show that the obliquity can be measured by adopting the χ2 minimization or the extended information criterion. We demonstrate the feasibility of our methodology by applying it to a multi-band photometry of a cloudless model Earth with future space missions such as the occulting ozone observatory (O3). We conclude that future space missions can estimate both the surface distribution and the obliquity at least for cloudless Earth-like planets within 5 pc.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Recent progress in observational techniques has revealed various physical properties of exoplanets beyond orbital parameters and planetary mass. Detections of atmospheric components have been reported for several systems using spectroscopy at the planetary transit and secondary eclipse (e.g., Charbonneau et al. 2002; Vidal-Madjar et al. 2003, 2004; Tinetti et al. 2007; Swain et al. 2008, 2009). Interior compositions can be inferred from planetary mass and radius (e.g., Léger et al. 2009; Charbonneau et al. 2009). Constructions of thermal maps of the planetary atmosphere have been proposed by Williams et al. (2006) and Cowan & Agol (2008). A longitudinal thermal map of HD 189733b has been constructed by Knutson et al. (2007) based on the method proposed by Cowan & Agol (2008).

Nevertheless, an identification of planetary surface components still remains an ambitious challenge. One of the promising approaches is to use the scattered light of exoplanets through the direct imaging observation (e.g., Seager et al. 2000; Ford et al. 2001; Sudarsky et al. 2005). Ford et al. (2001) focused on the inhomogeneity of the Earth's surface which causes diurnal variation of the scattered light. They computed the scattered light from a model Earth observed at a distance of 10 pc and showed that time variations of the scattered light in different photometric bands highly depend on the geological and biological features on the planetary surface such as ocean, land, and even vegetation. More detailed characterizations (including spectroscopy) of the scattered light of the Earth and its time variations are discussed via both Earth-shine observations (e.g., Woolf et al. 2002; Arnold et al. 2002; Montañés-Rodríguez et al. 2006) and simulations (Tinetti et al. 2006a, 2006b; Montañés-Rodríguez et al. 2006). These studies have suggested a future possibility of investigating the surface of Earth-like exoplanets by scattered light curves. We note that such time variations of scattered light are also applicable to determine the rotation period as shown by Pallé et al. (2008).

A variety of inversion techniques of the planetary surface from the scattered light curves have been proposed. Surprisingly, the first theoretical study of scattered light curves to make albedo maps was performed at the beginning of the twentieth century, although the authors assume asteroids and satellites for the target (Russell 1906). Cowan et al. (2009) performed principal component analysis (PCA) on multi-band photometric data of the Earth observed by the Extrasolar Planet Observation and Deep Impact Extended Investigation (EPOXI) mission and extracted spectral features which roughly correspond to land and ocean. They also checked the time variation of these components and translated it to the longitudinal distribution of these components based on the formulation by Cowan & Agol (2008). Oakley & Cash (2009) paid attention to the gap of reflectivity between ocean and land and reproduced a longitudinal map of land. Fujii et al. (2010, hereafter F10) have developed a methodology to estimate the areas of ocean, soil, vegetation, and snow from multi-band photometry and showed that the area of these components can be recovered from mock observations of a cloudless Earth.

Since these authors focused on diurnal variations in mapping the surface, the resultant maps have only longitudinal resolution with little information on the latitudinal distribution. One of the goals of the present paper is to develop a method to map inhomogeneous surfaces of exoplanets with both longitudinal and latitudinal resolutions using both diurnal and annual variations of scattered light.

We also consider the determination of the planetary obliquity from time variation of planetary light. The obliquity is an important property of Earth-like planets with its strong implications for climate, habitability (e.g., Williams & Kasting 1997; Williams & Pollard 2003), and planetary formation. N-body simulations of the final stage of terrestrial planet formation indicated that the distribution of the obliquity ζ is isotropic (Agnor et al. 1999; Chambers 2001; Kokubo & Ida 2007). Gaidos & Williams (2004) modeled the infrared light curves of exoplanets and showed how the obliquity affects an annual variation. Oakley & Cash (2009) also pointed out a possibility of determining a planet's obliquity. In this paper, we demonstrate that the obliquity is estimated simultaneously with the global map of the planet by analyzing the scattered light curves over its orbital period.

The rest of the paper is organized as follows. We first review the estimation method of the weighted area from multi-band photometry proposed by F10 and describe our methodology to reconstruct the planetary surface and the obliquity measurement in Section 2. Assuming a future satellite mission for the direct imaging of Earth-like planets, we apply our methodology to mock observations based on real data of the scattering properties of the Earth in Section 3. Finally, we summarize our results in Section 4.

2. METHODS

Let us briefly summarize the method of reconstructing a planetary surface from scattered light curves by F10. The scattered light from the planetary surface is characterized by the bidirectional reflectance distribution function (BRDF), f0, ϑ1, φ; λ)[str-1], where ϑ0, ϑ1, and φ are the incident zenith angle, the scattering zenith angle, and the angle between the incident and scattered light, respectively (Figure 1(a)). The BRDF is also a function of the wavelength λ. Using the BRDF at (ϕ, θ) fixed on the sphere (Figure 1(b)), the total scattered intensity at a given phase is provided by

Equation (1)

where F*(λ) is the incident flux at wavelength λ, Rp is a planetary radius, SVI is the surface area facing the observer and illuminated by the host star (we call it the visible and illuminated area), and dΩ = sin θdθdϕ. The position on the surface (ϕ, θ) fully specifies ϑ0, ϑ1, and φ, and thus we write f(ϕ, θ; λ) ≡ f0(ϕ, θ), ϑ1(ϕ, θ), φ(ϕ, θ); λ).

Figure 1.

Figure 1. Schematic configurations of the planetary surface and system. Panel (a) displays the definitions of the arguments of the BRDF f0, ϑ1, φ; λ): the incident zenith angle, ϑ0, the zenith angle of scattering, ϑ1, and the relative azimuthal angle between the incident and scattered light, φ. Panel (b) shows the spherical coordinate fixed on the planetary surface: the complementary latitude, θ, and the longitude, ϕ. The spin rotation axis is indicated by θ = 0. The spherical coordinate of the vector from the planetary center to observer is denoted by (ϕobs, θobs). Then, the spin rotation is described by Φ ≡ 2π − ϕobs. Panel (c) illustrates the coordinate system to specify the phase of the planet. The Θ denotes the orbital longitude measured from the summer solstice (i.e., the angle between the incident ray and the projected vector of the spin rotation axis to the orbital plane). The obliquity ζ is the angle between the spin rotation axis and a normal vector of the orbital plane. In this paper, we assume that the line of sight is perpendicular to the orbital plane.

Standard image High-resolution image

In order to make a fitting model for the multi-band photometric data, F10 classified the planetary surface into four types: ocean, snow, soil, and vegetation. Denoting the specific BRDF of each type k by fk0,  ϑ1,  φ), one can write the local BRDF at position (ϕ, θ) as a summation of the specific BRDFs of the kth-type surface weighted by local cover fractions m(k)(ϕ, θ):

Equation (2)

where NC is the number of surface types and NC = 4 in this case. Assuming that the scattering is isotropic (Lambertian), they approximate the BRDF by wavelength dependence of albedo ak(λ),

Equation (3)

Under the above assumption, Equation (1) reduces to

Equation (4)

Assuming that they know the spectra for each component, F10 developed a reconstruction method for the weighted area of the kth component,

Equation (5)

where W(ϕ, θ) is the weight function. The above definition reduces Equation (4) to a set of linear discrete equations:

Equation (6)

where C is a factor that depends on the phase angle, i.e., the planetocentric angle between the host star and the observer, and λb is the center wavelength of the bth band. Using a cloudless model Earth, they showed that the weighted area can be approximately recovered by solving Equation (6) with an inequality condition Ak>0, and the summation of Ak(t) is approximately constant over time, ∑Ak(t) ≈ const. ≈ R2p. It means that one can roughly estimate Rp as well by their method. Therefore, we use the fractional area Ak/R2p normalized by R2p in what follows. F10 have considered the spin rotation of a planet only and recovered the weighted area during one day of the planet and the weighted area can be converted to longitudinal maps by the inversion method proposed by Cowan & Agol (2008) and Cowan et al. (2009).

2.1. An Inverse Problem of a Two-dimensional Map

In the present paper, we consider the orbital motion to recover a two-dimensional world map of the planetary surface. We assume a model planet on a face-on circular orbit, which is the most promising case for planet mapping. We also assume that the orbital period is Porb = 365 day and the spin rotational period is Pspin = 24 hr. We denote the phases of the orbital motion and spin rotation by Θ and Φ (Figure 1(c)), respectively. Our mock observation is assumed to be performed over one year from Θ = ΘS to Θ = ΘS + 2π. There are two unknown parameters in this geometry: the obliquity ζ and the initial orbital longitude ΘS. We discuss the measurement of ζ and ΘS in Section 2.3.

For our conventional geometry, the weight function is given by

Equation (7)

We provide the explicit form of the visible and illuminated area, SVI in Appendix A. Note that the weight function is normalized to unity by integrating over the sphere:

Equation (8)

Figure 2 shows the behavior of the weight function W(ϕ, θ; Θ; Φ; ζ) for cases of ζ = 0°, 45°, and 90°. The weight function covers the region of 0° < θ < 90° + ζ and 0° < ϕ < 360° as the planet rotates around the host star and its spin axis. As a result, the scattered light curves contain information about the land distribution up to 50(1 + sin ζ)% of the total surface area in principle.

Figure 2.

Figure 2. Geometry dependence of the weight function W(ϕ, θ; Θ; Φ; ζ) in the visible and illuminated area SVI. Annual variations are indicated by different thin lines. Solid, large-dashed, middle-dashed, and small-dashed lines indicate Θ = 0°, 90°, 180°, and 270°, respectively. Because the weight function solely depends on ϕ + Φ, in other words, longitude minus the phase of the spin rotation, we adopt ϕ + Φ to x-axes instead of showing diurnal variations explicitly. Thick lines indicate the boundary of the visible area, θ = θV(ϕ; Φ; ζ) in Equation (A3). Each panel shows a different obliquity, ζ = 0° (panel a), 45° (panel b), and 90° (panel c). Inner and outer contours indicate W(ϕ, θ; Θ; Φ; ζ) = 0.6 and 0.3, respectively.

Standard image High-resolution image

We consider how to deduce the local cover fraction m(k)(ϕ, θ) if a set of the weighted area throughout the orbital and spin rotational periods is given. In order to solve this problem, we apply one of the techniques of tomography, the linear inverse problem, to the data. We assume that the planetary surface consists entirely of land and ocean and do not consider the other components nor further decompose them into clouds, vegetation, or soil in what follows. Therefore, we use the land cover fraction m(ϕ, θ) for the surface classification (k = land). The surface at (ϕ, θ) covered by land only (ocean only) is expressed by m(ϕ, θ) = 1 (m(ϕ, θ) = 0). We also denote the weighted area of land by A(Φ(ti), Θ(ti); ζ). In Section 3, we also consider the components of vegetation and soil using mock photometric data.

To begin with, let us discretize the observing time in N epochs and pixelize the planetary surface in M pixels. The weighted area recovered at the ith epoch ti is written as

Equation (9)

where sj indicates the jth pixel on the surface. The weighted cover fraction 〈mij is defined as

Equation (10)

Here, let us define the pixel-averaged cover fraction

Equation (11)

Under the assumption that the weighted cover fraction approximately equals the pixel-averaged cover fraction

Equation (12)

we obtain

Equation (13)

We introduce the data vector of weighted area ${\boldsymbol{a}}$, the design matrix G, and the model vector of cover fraction ${\boldsymbol{m}}$ as

Equation (14)

where σi is the error of the weighted area. Then we can rewrite Equation (13) as

Equation (15)

The above equation can be solved for ${\boldsymbol{m}}$ by minimizing χ2:

Equation (16)

under inequality constraints:

Equation (17)

where ${\boldsymbol{a}}_\mathrm{data}$ is the observed weighted area with noise. We denote ${\boldsymbol{m}}$ that minimizes Equation (16) by ${\boldsymbol{m}}_{\mathrm{est}}$.

2.2. Solving the Inverse Problem

We are now in a position to solve Equation (16) with condition (17) using an idealized data set. Since one can choose an arbitrary division for the surface modeling, we use the following pixelization:

Equation (18)

The above pixelization has a constant solid angle for a given ζ and the number of pixels does not change for different ζ's. We adopt Mϕ = 9 and Mθ = 23 (see the right panel of Figure 3 for the ζ = 90° case) and assume the numbers of epochs for diurnal variations NΦ = 23 and for annual variations NΘ = 23. In Appendix B, we discuss the dependence of model degeneracies on the pixelization.

Figure 3.

Figure 3. Input map (left) and reference map (right). The left panel shows the land distribution of the Earth based on the data set of the ISLSCP II fixed land/water masks (http://islscp2.sesda.com). Right panel indicates the pixel-averaged land cover fraction ($\overline{m}_j$) computed by averaging the left panel over pixels.

Standard image High-resolution image

We create the synthetic data of the weighted area using 1° × 1° fixed land/water masks of the ISLSCP II (input map, left panel of Figure 3). The pixel-averaged land cover fraction of this data (reference map) is shown in the right panel of Figure 3.

We fix the obliquity ζ = 90° and Θ(t0) = ΘS = 0° and compute the weighted area Ainput(Φ, Θ; ζ = 90°) from the input map (not from the reference map). We compute the data Adata by adding a Gaussian noise Ng to Ainput(Φ, Θ; ζ = 90°):

Equation (19)

We denote the standard deviation of the Gaussian random variable Ng by σ ≡ 〈N2g1/2.

The linear inverse problem (Equation (15)), or the equivalent least-square problem (Equation (16)), under the inequality constraints (Equation (17)) can be solved with the Bounded Variable Least-Squares Solver (BVLS) developed by Lawson & Hanson (1974, 1995).3 The BVLS is a generalization of the non-negative least square described in Lawson & Hanson (1974, 1995) and uses the QR decomposition (see also Menke 1989). The BVLS works for both overdetermined and underdetermined (ill-condition) problems. Even for ill-condition problems, the BVLS provides one of the (non-unique) solutions (Lawson & Hanson 1995).

By solving the least-square problem (Equation (16)) with the BVLS algorithm, we obtain the recovered map ${\boldsymbol{m}}_{\mathrm{est}}$. Figure 4 displays the two-dimensional images of the weighted area (left), the recovered maps (middle), and the prediction errors, $[A_\mathrm{data} (\Phi (t_i), \Theta (t_i)) - G {\boldsymbol{m}}_{\mathrm{est}}]/\sigma$ (right) for ζinput = 90°. In this figure, we regard ζ and ΘS as fixed parameters and use the input values ζ = ζinput = 90° and ΘS = ΘS,input = 0° for Equation (14). We consider four cases of the noise, σ = 0, 0.01〈Ainput〉, 0.1〈Ainput〉, and 0.3〈Ainput〉, where 〈Ainput〉 is the average of Ainput(ϕ, θ; ζ = 90°) and is approximately equal to 0.3 for the case of Earth. As discussed in Appendix C, we confirmed that the estimated map is the unique solution under the inequality constraint. Comparing the recovered maps with the reference map (right panel in Figure 3), we conclude that one can approximately recover the land distribution from the two-dimensional map of the weighted area.

Figure 4.

Figure 4. Two-dimensional images of the weighted area A(ϕ, θ) (left panels), recovered maps (middle panels) for ζ = 90°, and prediction errors, $[A_\mathrm{data} (\Phi (t_i), \Theta (t_i)) - G {\boldsymbol{m}}_{\mathrm{est}}]/\sigma$ for different noises, σ = 0, 0.01〈Ainput〉, 0.1〈Ainput〉, and 0.3〈Ainput〉 from top to bottom panels.

Standard image High-resolution image

We also solve the inverse problem in the case of the obliquity of Earth, ζinput = 23fdg45 (Figure 5). Even for a small obliquity like Earth's, one can recover the land distribution of 50 [1 + sin(ζ = 23fdg45)] ≈ 70 % of the planetary surface. We conclude that our methodology can recover main features of the continents on the planetary surface even if the planet has a small obliquity as is the case with Earth.

Figure 5.

Figure 5. Two-dimensional images of the weighted area A(ϕ, θ) (left panels) and recovered maps (right panels) for ζ = 23fdg45.

Standard image High-resolution image

2.3. Measurement of Obliquity

So far, we have assumed that the obliquity ζ and the initial phase ΘS in the orbit are known a priori. Next, we try to estimate the obliquity ζ by our method itself. Figure 6 indicates the obliquity dependence of images of the weighted area. We determine the best-fit value of ζ and ΘS by minimizing χ2 distance:

Equation (20)

where ${\boldsymbol{m}}_{\mathrm{est}}^{\zeta,\Theta _S}$ is determined by minimizing χ2 with given ζ and ΘS using the BVLS fitting. Figure 7 displays χ2(ζ, ΘS) for ζinput = 45° and ΘS,input = 180°. We use the amoeba routine (Press et al. 1992) to search for ζ and ΘS, which minimizes χ2(ζ, ΘS). We estimate errors of ζ and ΘS by the bootstrap resampling because errors are originated not only from additional Gaussian noise but also from pixelization, linearization, and other systematics. After 103 iterations of the bootstrap, we obtain ζest = 46fdg3 ± 0fdg6 and ΘS,est = 181fdg8 ±  0fdg8 for 10 % additional noise and ζest = 50fdg3 ± 2fdg3 and ΘS,est = 183fdg0 ± 2fdg3 for 30 % additional noise. The estimated values of both obliquity and initial phase in orbit agree remarkably well with the input values.

Figure 6.

Figure 6. Obliquity dependence of two-dimensional images of the weighted area.

Standard image High-resolution image
Figure 7.

Figure 7. χ2(ζ, ΘS) for ζinput = 45° and ΘS,input = 180° (cross) with 10% additional noise. The estimated obliquity and initial orbital longitude are ζest = 46fdg3 ± 0fdg6 and ΘS,est = 181fdg8 ± 0fdg8, respectively.

Standard image High-resolution image

3. APPLICATION TO MOCK OBSERVATION OF MULTI-BAND PHOTOMETRY

Now, we apply our methodology to a mock multi-band observation of a cloudless Earth as a more realistic demonstration. We simulate scattered light curves of a cloudless Earth in the same manner as the simulation in F10. We use the solar spectrum as an incident flux on the model Earth. We assume an observer seeing the planet on a face-on and circular orbit at a distance of 5 pc (Case A) or 10 pc (Case B) from it. The scattered light from land is computed using the BRDF of actual land surface assigned by the MODIS data set ("snow-free gap-filled MODIS BRDF Model Parameters"). We adopt the data of April and ignore the seasonal variation of the BRDF at each point on the planetary surface. The scattered light from ocean is calculated with the BRDF model for wavy ocean described in Nakajima & Tanaka (1983). In our model, the snow cover regions around the poles are replaced by ocean. We also include the effect of Rayleigh scattering by the atmosphere with a single scattering approximation between the atmosphere and the underlying surface. Neither the effect of clouds nor the molecular absorption is, however, incorporated.

Our assumed observation parameters are listed in Table 1. As an observing system, we assume a satellite telescope with 1.1 m aperture. This assumption comes from the basic architecture proposed for the occulting ozone observatory (O3) (Kasdin et al. 2010), which is a satellite mission for direct imaging of exoplanets in UV/optical/near-IR bands shading the light from the host star by a 30 m external occulter. The scattered light is computed at the central wavelength of each band and multiplied by its band width. We consider four bands centered at the wavelengths listed in Table 2, which correspond to the bands of the MODIS land BRDF data. We assume 0.1 μm as the bandwidth.

Table 1. Observation Parameters

Symbol Quantity Value Unit
texp Exposure time 24/23×3600 s
n Folded days 14 days
D Diameter of telescope aperture 1.1 m
Ψ Sharpness 0.0433  
α Pixel scale 0.03125 arcsec pixel−1
h End-to-end efficiency 0.5  
υ Dark rate 0.001 counts s−1
κ Read noise 2 $\sqrt{{\rm counts}}$/read
QE Quantum efficiency 0.91  
Ωz Zodiacal light in magnitude 23 mag arcsec2
Z Zodiacal light 1  
F0 Zero flux 1.4 × 104 (band 1) cts cm-2 nm s−1
    9.6 × 103(band 2) cts cm-2 nm s−1
    7.2 × 103(band 3) cts cm-2 nm s−1
    4.1 × 103(band 4) cts cm-2 nm s−1

Download table as:  ASCIITypeset image

Table 2. Band Definition

Band Wavelength Band Width
  λb (μm) Δλ (μm)
1 0.469 0.1
2 0.555 0.1
3 0.645 0.1
4 0.8585 0.1

Download table as:  ASCIITypeset image

We divide the orbital phase into 23 (Θi = 2πi/23,  i = 0, 1,..., 22) and the spin rotational phase into 23 (Φi = 2πi/23,  i = 0, 1,..., 22). We compute the scattered light curves in four bands with the exposure time texp = 24/23 hr throughout one year. Then, we stack data for 14 days centered at each orbital phase Θi and fold the light curves according to its spin rotation period to obtain diurnal curves at the orbital phase, assuming that the spin rotational period is already known through periodogram analysis (Pallé et al. 2008). Thus, the resultant light curves have 23 × 23 data points for each band.

As for the noise, we consider the photon shot noise, the exzodiacal light, dark noise, and read noise. We ignore the leakage of the light from the host star. Thus, the signal-to-noise ratio (S/N) is expressed as

Equation (21)

where S, Nz, Nd, and Nr are the scattered light from the planet, and the contribution from zodiacal light, dark noise, and read noise, respectively. Specifically, they are given as

Equation (22)

Equation (23)

Equation (24)

The parameters and their values we adopt are listed in Table 1 (see, e.g., Savransky et al. 2010). The zero flux F0 for calibration is calculated by linearly interpolating the spectrum of Vega given by Tables 1 and 2 in Colina et al. (1996) and shown in Table 3 in this paper. In order to quantitatively assess each noise level, we list the amplitudes of these noises per epoch in Table 3 together with the total average of the signal S in the case of ζ = 90°. The averaged S/N per epoch is also listed in Table 3. We note that the averaged signal per epoch is comparable with the noise level.

Table 3. Assumed Signal and Noises in Four Bands

Band b 1 2 3 4
Averaged signal $\bar{S}$ (case A; 101 cts epoch−1) 40.9 40.3 36.9 40.9
Averaged signal $\bar{S}$ (case B; 101 cts epoch−1) 10.2 10.1 9.2 10.2
Exozodi Nz (103 cts epoch−1) 12.4 9.2 7.0 4.0
Dark noise Nd (103 cts epoch−1) 1.2
Read noise Nd (103 cts epoch−1) 1.3
Averaged S/N (case A; 1 epoch−1)a 3.3 3.6 3.7 4.9
Averaged S/N (case B; 1 epoch−1)a 0.83 0.93 0.94 1.3

Note. aThe averaged S/Ns are defined by $\bar{S}/\sqrt{\bar{S}+N_z+N_d+N_r}$.

Download table as:  ASCIITypeset image

The obtained mock data Ib) at each epoch (Θ(ti), Φ(ti)) is decomposed into the weighted area of four surface types, (ocean, vegetation, soil, and snow) by solving Equation (6) following F10. The weighted area of land is computed by summing up that of soil, vegetation, and snow, that is, A(Θ, Φ) = Asoil(Θ, Φ) + Avegetation(Θ, Φ) + Asnow(Θ, Φ). While F10 used nonlinear fitting in order to constrain Ak to be positive, we use the BVLS linear solver instead without upper constraints. The resultant A(Θ, Φ) as a function of Θ and Φ is displayed in the left panels of Figure 8.

Figure 8.

Figure 8. Simulated weighted area from mock photometric observation (left), its recovered map (middle), and prediction error (right). Top and bottom panels indicates cases A and B, respectively. We note that the snow component such as the South Pole in the input data is replaced by ocean.

Standard image High-resolution image

We now apply the reconstruction method described in Section 2 (see Equations (14), (16), and (17)) to the set of obtained A(Θ, Φ). For the errors σi in Equation (14), we first compute the variance of the weighted area by 100 times resampling of photon counts according to Poisson statistics, and we call such variance as σ2i,res. The σi,res, however, turned out not to relevant for σi in Equation (14) because it often results in zero or very small errors when Ai = 0 due to the boundary condition (Ai ⩾ 0). Therefore, we substitute an average value of errors σ = ∑Ni=1σi,res/N to Equation (14) instead. For the case that all estimated errors of Ak are positive, the results with σi = σi,res do not change significantly. In the fitting, we assume that the local cover fraction of each surface type remains unchanged and thus ignore the seasonal variations of vegetation, soil, and snow. The center and right rows of Figure 8 display recovered maps and prediction errors for ζinput = 90° of the cases A and B. The major features of the surface can be seen in the resultant maps. The averaged prediction errors $ |A_\mathrm{data} (\Phi (t_i), \Theta (t_i)) - G {\boldsymbol{m}}_{\mathrm{est}}|/N$ are 0.53 〈Adata〉 (case A) and 0.85 〈Adata〉 (case B), respectively.

The land and ocean distributions have been considered so far. Next, we create color composition maps of vegetation and soil on the land and ocean recovered map (Figure 9). Adopting the same method of the land recovery to the weighted areas of soil (Asoil(ti)) and vegetation (Avegetation(ti)), we derive the soil and vegetation distributions and blend them by yellow (soil) and green (vegetation) over Figure 8, in other words, the land (white) and ocean (blue) distributions. Even using the weighted areas of soil and vegetation separately for the recovery, the resulting vegetation and soil maps almost distribute over the land recovered with the summation of the weighted areas (Aland(ti) = Asoil(ti) + Avegetation(ti) + Asnow(ti)). In this figure, snow is ignored because the contribution of it is significantly small. The concentrations of vegetation at South America and soil at middle Africa seen in case A indeed correspond to the Amazon forest and the Sahara desert, respectively.

Figure 9.

Figure 9. Recovered color maps of soil and vegetation from the mock observations. Soil and vegetation are blended by yellow and green on the land recovered maps shown in the middle panels of Figure 8. Top left and right panels indicate case A and case B, respectively. Snow is replaced by ocean in this simulation. Bottom panel is a classification map of ocean (blue), snow (pink), vegetation (green), and soil (yellow) based on the IGBP classification. Our classification of soil and vegetation is listed in Table 2 of F10.

Standard image High-resolution image

Now, we try to measure the obliquity for the mock observation. Figures 1013 display the light curves of four bands with ΘS,input = 180°, corresponding to different obliquities, ζinput = 90°, 60°, 45°, and 30°, respectively. The noises assumed in these figures are computed in case A. Performing the χ2 fitting described in Section 2.3, we estimate the best-fit obliquities and the initial orbital longitude. The predicted curves with the best-fit values are shown by solid lines in Figures 1013. The input and the best-fit values and the reduced χ2 are listed in Table 4. We compute the χ2 for the land distribution (Equation (20)). Therefore, the degree of freedom is 23 × 23 − 23 × 9 = 322. It is likely that the inequality constraints and the other systematics such as the Lambert assumption induce the relatively high reduced χ2 values (∼2). One can see both diurnal and annual variations of light curves in these figures. The larger variations are seen in band 4 (0.8585 μm) due to large albedos of the land component (soil and vegetation; see Figure 7 in F10). There is "pinching" of the light curve at ΘS = 0° (i ∼ 264) for ζinput = π/4 because the maximum of the weight function is located at the planet's pole. In this period, the physical position of the maximum stays at the pole for the daily motion. As a result, there are little variations in the period. However, the "pinching" seen at ΘS ∼ 180°(i ∼ 0) for ζinput = 90° has a different origin. In this period, the weight function moves around the Southern Hemisphere, that is, ocean mainly.

Figure 10.

Figure 10. Mock light curves in units of reflectivity for ζinput = 90°. Each panel indicates a different band: top left, top right, bottom left, and bottom right panels correspond to bands 1 (0.469 μm), 2 (0.555 μm), 3 (0.645 μm), and 4 (0.8585 μm). Dashed vertical lines divide each epoch Φ, Θ stack in 14 days observations. There are 23 data points in one epoch and 23 epochs in the whole data set. Therefore, there are 23 = 529 data points in the whole data set (i = 1, 2,..., 529). Wider versions of two shaded regions labeled in (a) and (b) are inserted in the mini panels. The predicted curves with the best-fit obliquities ζest = 89fdg8 and initial longitude ΘS,est = 183.8 are drawn by solid lines. The observational noises are computed on the assumption of case A.

Standard image High-resolution image
Figure 11.

Figure 11. Same as Figure 10 for ζinput = 60°.

Standard image High-resolution image
Figure 12.

Figure 12. Same as Figure 10 for ζinput = 45°.

Standard image High-resolution image
Figure 13.

Figure 13. Same as Figure 10 for ζinput = 30°.

Standard image High-resolution image

Table 4. Estimated Obliquity and the Initial Orbital Longitude by χ2 Minimization and the EIC

Case ζinput ΘS,input The χ2 Fit The χ2 Fit χ2 χ2/dof EIC EIC
      ζest ΘS,est     ζest ΘS,est
A 90° 180° 89fdg8 ± 1fdg2 183fdg8 ± 2fdg2 650.3 2.02 89fdg0 184fdg5
A 60° 180° 64fdg7 ± 4fdg8 175fdg3 ± 3fdg5 661.2 2.05 65fdg3 176fdg5
A 45° 180° 44fdg4 ± 3fdg3 179fdg8 ± 4fdg1 731.1 2.27 44fdg5 177fdg5
A 30° 180° 32fdg5 ± 3fdg0 188fdg6 ± 5fdg6 677.5 2.10 30fdg7 190fdg5
B 90° 180° 89fdg8 ± 3fdg5 190fdg8 ± 17fdg5 596.1 1.85 88fdg1 187fdg5
B 60° 180° 79fdg7 ± 8fdg1 171fdg8 ± 24fdg8 568.5 1.77 48fdg5 155fdg6
B 45° 180° 38fdg2 ± 21fdg6 231fdg8 ± 43fdg6 609.5 1.89 26fdg7 132fdg6
B 30° 180° 89fdg2 ± 17fdg1 210fdg3 ± 32fdg7 714.4 2.23 7fdg9 138fdg6

Download table as:  ASCIITypeset image

The χ2 maps for the four input obliquities are displayed in Figure 14 and the best-fit obliquities and initial longitudes are listed in Table 4. For case A, the χ2 fitting provides the estimated values close to the input ones within 10°. Figure 15 shows the difference of the predicted light curves with the best-fit obliquity ζest = 44.4 and that shifted ±3fdg3(∼1σ level estimated by the bootstrap resampling), ζ = 47fdg7 and 41fdg1.

Figure 14.

Figure 14. ζ and ΘS dependence of χ2(ζ, ΘS) of the mock observational data set (case A) for ΘS,input = 180° and the different input obliquities, ζinput = 90° (top left), ζinput = 60° (top right), ζinput = 45° (bottom left), and ζinput = 30° (bottom right). The input values are indicated by crosses. The estimated obliquity and initial orbital longitude are listed in Table 4.

Standard image High-resolution image
Figure 15.

Figure 15. Difference between the light curve with the best-fit values (ζest = 44fdg4, ΘS,est = 179fdg8), with a 3fdg3 shifted obliquity (ζ = 47fdg7, ΘS = 179fdg8; solid curves) and with a −3fdg3 shifted obliquity (ζ = 41fdg1, ΘS = 179fdg8; dashed curves).

Standard image High-resolution image

The results for case B are also listed in Table 4. Figure 16 displays the χ2 map for case B. The estimated values for ζinput = 30° and 60° tend to deviate from the input one. As shown in the bottom right panel, the result for ζinput = 30° is significantly biased. It is likely that the systematics affect the obliquity estimates for this noise level. We also perform the other method called the extended information criterion (EIC) in Appendix D. Although the tendency that the estimated value is biased to ζ = 90° vanishes, the best-fit value does not improve significantly.

Figure 16.

Figure 16. Same as Figure 14 for case B.

Standard image High-resolution image

In short, the planetary obliquity can be estimated well for case A, while, for case B which has larger noises, it is marginal to derive reliable values of the estimated obliquity by our method.

4. SUMMARY AND DISCUSSION

We have developed a method of reconstructing the two-dimensional planetary surface via diurnal and annual variation of the scattered light. Applying the method to the mock photometric data, we have demonstrated that our method works for the mock Earth model, while this model has a lot of simplifying assumptions as follows: (1) cloudlessness, (2) a face-on circular orbit, (3) known reflectance spectra, (4) lack of atmospheric absorption, (5) known rotation rate, (6) a static map, and (7) the absence of a moon. We also found that the planetary obliquity can be estimated by this method. With our method, future satellite missions such as the occulting ozone observatory (Kasdin et al. 2010) might provide "a global map" of Earth-like exoplanets. While only the terrestrial planets have been considered in this paper, our method might be applicable to any planets with an inhomogeneous surface, including Jupiter-like exoplanets.

In this paper, we ignored the effect of clouds and expected that clouds affect the estimation like a statistical noise because of the relatively short time variation of clouds. The PCA performed by Cowan et al. (2009) is a promising approach because they could separate the land and ocean compositions even though they used the EPOXI data that contain the cloud effect. The effect of clouds is discussed elsewhere (Y. Fujii et al. 2010, in preparation). We also assumed a face-on circular orbit in this paper. This assumption might be too severe for practical applications. We will generalize our method in the next paper.

We are deeply grateful to Yasushi Suto, Atsushi Taruya, and Edwin L. Turner for helpful and insightful discussion. We are thankful to Dmitry Savransky, Jeremy N. Kasdin, and David N. Spergel, who enlightened us about the observing instruments designed for future satellite missions. We also thank the referee, Nicolas Cowan, for a lot of constructive comments. H.K. is supported by a Japan Society for Promotion of Science (JSPS) Grant-in-Aid for science fellows. This work is also supported by a Grant-in-Aid for Scientific Research from JSPS and from the Japanese Ministry of Education, Culture, Sports, Science, and Technology (20·10466 and 22·5467), and by the JSPS Core-to-Core Program "International Research Network for Dark Energy."

APPENDIX A: VISIBLE AND ILLUMINATED AREA

The visible and illuminated area is expressed as

Equation (A1)

The boundary lines of visible area θV(ϕ; Φ; ζ) and illuminated area θI,+(ϕ; Θ; Φ; ζ) and θI,−(ϕ; Θ; Φ; ζ) are obtained by solving the following equation:

Equation (A2)

The explicit equations of θV(ϕ; Φ; ζ), θI,+(ϕ; Θ; Φ; ζ), and θI,−(ϕ; Θ; Φ; ζ) are expressed as

Equation (A3)

Equation (A4)

Equation (A5)

where

APPENDIX B: PIXELIZATION OF PLANETARY SURFACE

We note here that the design matrix alone does not fully specify ${\boldsymbol{m}}_{\mathrm{est}}$ without any constraints. We perform singular value decomposition (SVD) of the design matrix:

Equation (B1)

where U and V are N × N and M × M unitary matrices, and Λ is a N × M matrix:

Equation (B2)

Equation (B3)

with lj being the singular value of G in descending order. We denote the jth row of V by the vector ${\boldsymbol{v}_j}$.

In general, there are Nnull zero components in the singular values, $(l_{M-N_\mathrm{null}+1}=l_{M-N_\mathrm{null}+2}=l_{M}=0)$. Corresponding ${\boldsymbol{v}}$ vectors are called null vectors. Linear combination of the null vectors does not affect the data vector,

Equation (B4)

where βnull,s is an arbitrary coefficient. Equation (B4) indicates that the predicted model has a freedom to add any linear combination of the null vectors.

The solid curve in the left panel of Figure 17 displays the singular values in descending order (this plot is referred to as spectrum of data kernel) in the case of Mϕ = NΦ = NΘ = 23 and Nθ = 9. Due to the truncation errors, the singular values obtained by numerical algorithms of the SVD do not vanish exactly. However, only three of them (j = M − 2, M − 1, and M) have significantly smaller values than other diagonal elements. We regard the elements smaller than 10−4 as zero and the corresponding row vectors of V as null vectors.

Figure 17.

Figure 17. Singular values of the design matrix G in descending order (left panel). Dotted, solid, and dashed curves correspond to Mϕ = 22, Mϕ = 23, and Mϕ = 24, respectively. All curves assume Mθ = 9. Right panel shows the number of vectors with singular values below 10−4 as a function of Mθ and Mϕ = NΦ = NΘ.

Standard image High-resolution image

Although the presence of null vectors does not directly lead to the indefiniteness if one adopts the inequality constraints (Equation (17)), a smaller number of null vectors makes it easier to interpret an uncertainty of the recovered map. Therefore, it is useful to know how the number of null vectors depends on pixelization. Then, one can decrease the number of the null vectors by an appropriate choice of pixelization. For simplicity, we fix Mϕ = NΦ = NΘ and estimate the number of the null vectors by changing Mθ and Mϕ. The left panel in Figure 17 shows the spectra of the singular value in descending order for three cases; Mϕ = 22 (dotted), Mϕ = 23 (solid), and Mϕ = 24 (dashed) with Mθ = 9. The case of Mϕ = 23 has a steeper spectrum than the others. It means that the choice of Mϕ = 23 makes it easier to identify null vectors. The right panel of Figure 17 indicates the number of null vectors for different sets of Mϕ and Mθ, where we define vectors with singular value below 10−4 as null vectors. There is a general tendency that the odd number bin of Mϕ has lesser Nnull. Although the selection of smaller values of Mθ or Mϕ generally leads to a smaller number of the null vectors, it degrades the resolution of the map. As a compromise, we decided to adopt Mϕ = 23 and Mθ = 9 for our fiducial values of the model in this paper. Figure 18 indicates the (unit) null vectors for Mϕ = 23 and Mθ = 9, ${\boldsymbol{v}}^{\mathrm{null},1} = {\boldsymbol{v}}_{M-2}, {\boldsymbol{v}}^{\mathrm{null},2} = {\boldsymbol{v}}_{{M-1}}$, and ${\boldsymbol{v}}^{\mathrm{null},3} = {\boldsymbol{v}}_{M}$, where ${\boldsymbol{v}}_j$ is the jth row of V. These three null vectors for Mϕ = 23 and Mθ = 9 affect the large-scale structure of the map but do not change the small-scale distribution like the continents of the present Earth.

Figure 18.

Figure 18. Three unit null vectors of our fiducial model Mϕ = NΦ = NΘ = 23 and Mθ = 9. These maps are drawn using the Hammer–Aitoff projection.

Standard image High-resolution image

APPENDIX C: UNIQUENESS OF THE RECOVERED MAP

In this appendix, we examine whether there is the freedom for adding any components of null vectors to the recovered map derived by the BVLS. Because of the boundary condition for ${\boldsymbol{m}}_{\mathrm{est}}$ (Equation (17)), the estimated values of mest j are classified into three types: (1) mest j = 1 (on the upper boundary), (2) mest j = 0 (on the lower boundary), and (3) 0<mest j < 1 (in between). We focus on types (1) and (2) to consider the uniqueness of our recovered map. We denote the index j of type (1) by j− and that of type (2) by j+. In order not to violate the boundary condition, the linear combination of null vectors should satisfy the following constraints:

Equation (C1)

Equation (C2)

These inequalities can be solved in terms of βnull,1 using the fact that vj is positive (see Figure 18):

Equation (C3)

Equation (C4)

Then, the following inequality is required so that βnull,1 exists

Equation (C5)

This equation can be reduced to the following expression:

Equation (C6)

where p(j+,j−) and b are two-dimensional vectors:

Equation (C7)

Equation (C8)

Now, let us denote the arguments of p and b by θα(j +, j−) and θβ, respectively, i.e.,

Equation (C9)

Equation (C10)

If |b| ≠ 0, the solution of Equation (C6) should satisfy

Equation (C11)

We calculate the above constraints for any combination of (j +, j−) and found that there is no θβ that satisfies Equation (C11) for all combinations of (j +, j−). Thus, the only solution allowed is βnull,2 = βnull,3 = 0. From equations (C3) and (C4), it follows that βnull,1 = 0. As a result, we confirmed that there is no room for adding extra components of null vectors to the recovered map.

APPENDIX D: ANALYSIS WITH THE EXTENDED INFORMATION CRITERION

In this appendix, we try to perform the model selection of ζest and ΘS,est using the EIC (e.g., Efron 1983; Konishi & Kitagawa 1996; Ishiguro et al. 1997) based on the Kullback–Leibler divergence. Although the model selection for ζ and ΘS based on the χ2 minimization is very convenient, the result might be biased for large noises compared with signal. The EIC is a bootstrap-based extension of Akaike Information Criterion defined as

Equation (D1)

where $ \mathcal {F}({\boldsymbol{a}}|{\boldsymbol{m}}_{\mathrm{est}}, \zeta _{\mathrm{est}}, \Theta _{S,\mathrm{est}})$ indicates the likelihood function. The best model or parameter set is given by minimizing the EIC. The bootstrap estimate of the bias bB is given by Ishiguro et al. (1997):

Equation (D2)

where $ {\boldsymbol{a}}^{\ast,k}, {\boldsymbol{m}}_{\mathrm{est}}^{\ast,k}, \zeta _{\mathrm{est}}^{\ast,k}$, and Θ*,kS,est indicate the kth bootstrap resample, ${\boldsymbol{m}}_{\mathrm{est}}$, ζest, and ΘS,est, and NEIC is the number of trials of the bootstrap. Assuming a Gaussian distribution for ${\boldsymbol{a}}$ with σ = 1, we derive

Equation (D3)

We adopt NEIC = 500 for bootstrap resampling. Figure 19 displays the dependence of the EIC on ζest and ΘS,est. The minimum EIC estimate listed in Table 4 is derived by a stepwise search at 1° interval.

Figure 19.

Figure 19. ζ and ΘS dependence of the EIC of the mock observational data set (case B) for ΘS,input = 180° and the different input obliquities, ζinput = 90° (top left), ζinput = 60° (top right), ζinput = 45° (bottom left), and ζinput = 30° (bottom right). The input values are indicated by crosses.

Standard image High-resolution image

The minimum EIC estimate listed in Table 4 is derived by a stepwise search at 1° interval. For case A, the EIC provides almost the same values as listed in Table 4. Figure 19 displays the EIC map for case B. Although the estimated obliquities for ζinput = 30°, 45°, and 60° by the EIC do not agree well with the input value (Table 4), the bias of ζinput = 60° and 30° seems to be improved compared to that from the χ2 fitting.

Footnotes

Please wait… references are loading.
10.1088/0004-637X/720/2/1333