Evidence for Nightside Water Emission Found in Transit of Ultrahot Jupiter WASP-33b

Yuanheng Yang CAS Key Laboratory of Planetary Sciences, Purple Mountain Observatory, Chinese Academy of Sciences, Nanjing 210023, China School of Astronomy and Space Science, University of Science and Technology of China, Hefei 230026, China Guo Chen CAS Key Laboratory of Planetary Sciences, Purple Mountain Observatory, Chinese Academy of Sciences, Nanjing 210023, China CAS Center for Excellence in Comparative Planetology, Hefei 230026, China Fei Yan Department of Astronomy, University of Science and Technology of China, Hefei 230026, China Xianyu Tan Tsung-Dao Lee Institute & School of Physics and Astronomy, Shanghai Jiao Tong University, Shanghai 201210, China Jianghui Ji CAS Key Laboratory of Planetary Sciences, Purple Mountain Observatory, Chinese Academy of Sciences, Nanjing 210023, China CAS Center for Excellence in Comparative Planetology, Hefei 230026, China
Abstract

To date, the dayside thermal structure of ultrahot Jupiters (UHJs) is generally considered to be inverted, but their nightside thermal structure has been less explored. Here we explore the impact of nightside thermal emission on high-resolution infrared transmission spectroscopy, which should not be neglected, especially for UHJs. We present a general equation for the high-resolution transmission spectrum that includes planetary nightside thermal emission. This provides a new way to infer the thermal structure of the planetary nightside with high-resolution transmission spectroscopy. Using the cross-correlation technique, we find evidence for the presence of an H2O emission signature on the UHJ WASP-33b during the transit, indicating an inverted temperature structure on its nightside. Such a result suggests a stronger heat transport through the circulation than currently expected. An alternative explanation is that the rotating visible hemisphere during transit leads to the potential contribution of the limb and dayside atmospheres to the detected emission signature. In the future, the combination of high-resolution full-phase curve spectroscopic observations and general circulation models will hopefully solve this puzzle and provide a complete picture of the three-dimensional nature of the chemistry, circulation, and thermal structure of UHJs.

journal: ApJL

1 Introduction

Ultrahot Jupiters (UHJs) are gas giant planets that orbit very close to their host stars with high dayside temperatures (Tdaysubscript𝑇dayT_{\mathrm{day}}italic_T start_POSTSUBSCRIPT roman_day end_POSTSUBSCRIPT \geq 2200 K, Parmentier et al., 2018). The intense stellar irradiation on the dayside of UHJs can lead to thermal dissociation of key radiatively active molecules, which are then transported to the nightside for recombination. This process results not only in global chemical inhomogeneities but also in significant changes in the radiative balance, leading to remarkable alterations in the dynamical pattern and thermal structure. Due to their observational accessibility, UHJs serve as an ideal laboratory for testing planetary atmospheric physics and models under extreme conditions.

General circulation models (GCMs) play a crucial role in deciphering the chemical and physical processes within the three-dimensional (3D) atmospheres of UHJs. In particular, they provide insights into the dynamics and temperature-pressure (T-P) structures. The day-night temperature differences in hot Jupiters could be controlled by various factors, including stellar irradiation, circulation patterns, radiative heating/cooling, and frictional drag (Showman & Guillot, 2002; Komacek & Showman, 2016; Tan & Komacek, 2019). In addition to the day-night temperature differences, the vertical temperature structures also vary from the dayside to the nightside. Typically, the temperature decreases with decreasing pressure (i.e., non-inverted T-P). However, several studies (Hubeny et al., 2003; Fortney et al., 2005, 2008) proposed that the T-P of planetary atmospheres strongly irradiated by stars bifurcates, especially in the presence of optical absorbers such as TiO and VO, resulting in a thermal inversion, i.e., the temperature increases with decreasing pressure. Parmentier et al. (2015) further presented that the non-gray thermal effects, which reduce the cooling capacity of the upper atmosphere, are the reason for the thermal inversion. Furthermore, metal atoms, molecules (such as silicon oxide and metal hydrides), and H- can also induce dayside temperature inversions in extremely irradiated hot Jupiters (Lothringer et al., 2018; Arcangeli et al., 2018; Parmentier et al., 2018). Gandhi & Madhusudhan (2019) showed that some refractory species other than TiO and VO or low infrared opacity can also cause thermal inversion. While current theories satisfactorily explain the dayside temperature inversion, there remains a lack of theory and observational constraints on the vertical temperature structure of the nightside atmosphere, which is empirically assumed to be either non-inverted or isothermal.

In recent years, transmission spectroscopy has revealed an increasing number of chemical species in exoplanet atmospheres, ranging from neutral or weakly ionized atoms to molecules. It typically assumes a dark planet transiting the host star and considers only the chromatic absorption of stellar light by planetary atmospheres. Several studies (Kipping & Tinetti, 2010; Chakrabarty & Sengupta, 2020; Martin-Lagarde et al., 2020; Morello et al., 2021) have attempted to investigate the effect of the nightside emission from hot Jupiters on the transit light curves and found that the chromatic dilution of the mid-infrared transit depth can be as large as 10-4, which may bias the inference of chemical abundances and thermal structures of some exoplanet atmospheres in the era of the James Webb Space Telescope. However, the effect of velocity-resolved nightside emission on transmission spectroscopy remains unexplored.

In this letter, we focus on the implications for high-resolution infrared transmission spectroscopy of wavelength-dependent thermal emission arising from diverse temperature structures on the nightside of UHJs. We take WASP-33b as an example for illustration, which is a UHJ with a mass of 2.2 MJsubscript𝑀JM_{\mathrm{J}}italic_M start_POSTSUBSCRIPT roman_J end_POSTSUBSCRIPT and a radius of 1.5 RJsubscript𝑅JR_{\mathrm{J}}italic_R start_POSTSUBSCRIPT roman_J end_POSTSUBSCRIPT orbiting a δ𝛿\deltaitalic_δ-Scuti A-type star with a period of 1.22 days (Collier Cameron et al., 2010). Despite being one of the most intensely irradiated hot Jupiters, WASP-33b exhibits heat transport efficiency comparable to other hot Jupiters (Zhang et al., 2018). Instead of limiting ourselves to the nightside thermal structures predicted by the current mainstream GCMs, we infer possible nightside thermal structures directly from the observational data and try to understand them in order to contribute to the theoretical development of 3D circulation patterns and energy transport efficiencies in UHJs.

2 Radiation Model of the Nightside Atmosphere of Planets

2.1 Planet’s Thermal Emission Signal vs. Transmission Absorption Signal

The planetary atmosphere can absorb light from the host star during transit. Each atomic or molecular species has a different extinction coefficient for the stellar spectrum at a given wavelength. Therefore, the information about the species of the planetary atmosphere is imprinted on the outgoing stellar spectrum during the transit. The classical transmission spectrum (Tmsubscript𝑇mT_{\mathrm{m}}italic_T start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT) is defined as the spectral flux ratio between out-of-transit and in-transit, given by (omitting the notation λ𝜆\mathrm{\lambda}italic_λ),

Tm=1FoutFinFout=1Reff2ψm(u)Rs2,subscript𝑇m1subscript𝐹outsubscript𝐹insubscript𝐹out1superscriptsubscript𝑅eff2subscript𝜓m𝑢superscriptsubscript𝑅s2\displaystyle T_{\mathrm{m}}=1-\frac{F_{\mathrm{out}}-F_{\mathrm{in}}}{F_{% \mathrm{out}}}=1-\frac{R_{\mathrm{eff}}^{2}\cdot\psi_{\mathrm{m}}(u)}{R_{% \mathrm{s}}^{2}},italic_T start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT = 1 - divide start_ARG italic_F start_POSTSUBSCRIPT roman_out end_POSTSUBSCRIPT - italic_F start_POSTSUBSCRIPT roman_in end_POSTSUBSCRIPT end_ARG start_ARG italic_F start_POSTSUBSCRIPT roman_out end_POSTSUBSCRIPT end_ARG = 1 - divide start_ARG italic_R start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⋅ italic_ψ start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT ( italic_u ) end_ARG start_ARG italic_R start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ,

where Reffsubscript𝑅effR_{\mathrm{eff}}italic_R start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT is the effective planetary radius as a function of wavelength, Rssubscript𝑅sR_{\mathrm{s}}italic_R start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT is the stellar radius, and ψmsubscript𝜓m\psi_{\mathrm{m}}italic_ψ start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT is the planetary rotational broadening profile for the transmission spectrum. We computed the rotational broadening kernel of the transmission spectrum according to the method of Boucher et al. (2023) and modified the code developed by Carvalho & Johns-Krull (2023) to implement fast rotational broadening over a wide range of wavelengths. We initially set the thickness of the atmosphere to z=5H𝑧5𝐻z=5Hitalic_z = 5 italic_H, where H𝐻Hitalic_H is the atmospheric scale height and tidal locking is assumed.

In addition to absorbing starlight from their host stars, planetary atmospheres also produce their own thermal emission. As the thermal emission travels from the inner to the outer atmosphere, it interacts with local matter. The spectral line features in the outgoing spectrum are determined by the vertical T-P profile and the chemical abundance profiles of the planetary atmosphere. The thermal emission spectrum (Tesubscript𝑇eT_{\mathrm{e}}italic_T start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT) can be expressed as (omitting the notation λ𝜆\mathrm{\lambda}italic_λ),

Te=FpRp2ψe(u)FsRs2,subscript����esubscript𝐹psuperscriptsubscript𝑅p2subscript𝜓e𝑢subscript𝐹ssuperscriptsubscript𝑅s2\displaystyle T_{\mathrm{e}}=\frac{F_{\mathrm{p}}\cdot R_{\mathrm{p}}^{2}\cdot% \psi_{\mathrm{e}}(u)}{F_{\mathrm{s}}\cdot R_{\mathrm{s}}^{2}},italic_T start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT = divide start_ARG italic_F start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT ⋅ italic_R start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⋅ italic_ψ start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT ( italic_u ) end_ARG start_ARG italic_F start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT ⋅ italic_R start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ,

where Fpsubscript𝐹pF_{\mathrm{p}}italic_F start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT is the thermal emission flux at the top of planetary atmosphere, Fssubscript𝐹sF_{\mathrm{s}}italic_F start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT is the stellar flux, Rpsubscript𝑅pR_{\mathrm{p}}italic_R start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT is the planetary radius, Rssubscript𝑅sR_{\mathrm{s}}italic_R start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT is the stellar radius, and ψesubscript𝜓e\psi_{\mathrm{e}}italic_ψ start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT is the rotational broadening profile for the thermal emission. Unlike the transmission spectrum, here we use the code from Carvalho & Johns-Krull (2023) directly to implement fast rotational broadening.

For high-resolution observations, we focus on the strength and profile of the spectral lines, which are closely related to local temperatures and temperature gradients in the planetary atmosphere. For planets with possible nightside temperature gradients, the spectral line features of the transmission spectrum are coupled with the spectral line features of the planet’s own thermal emission during transit. For WASP-33b, the magnitude of the line intensity of the Tmsubscript𝑇mT_{\mathrm{m}}italic_T start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT is about 104superscript10410^{-4}10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT. If there is a sufficient temperature gradient, the magnitude of the line intensity of the Tesubscript𝑇eT_{\mathrm{e}}italic_T start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT is also about 104superscript10410^{-4}10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT. Therefore, the signals from high-resolution transmission lines and thermal emission lines may be comparable.

2.2 General High-Resolution Transmission Spectrum with Planetary Nighside Emission

A transmission spectrum is defined as the spectral flux difference between out-of-transit and in-transit:

Δλ(t)¯λ,outλ,in(t)¯λ,out,subscriptΔ𝜆𝑡subscript¯𝜆outsubscript𝜆in𝑡subscript¯𝜆out\Delta_{\lambda}(t)\equiv\frac{\overline{\mathcal{F}}_{\mathrm{\lambda,out}}-% \mathcal{F}_{\mathrm{\lambda,in}}(t)}{\overline{\mathcal{F}}_{\mathrm{\lambda,% out}}},roman_Δ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_t ) ≡ divide start_ARG over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_λ , roman_out end_POSTSUBSCRIPT - caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_in end_POSTSUBSCRIPT ( italic_t ) end_ARG start_ARG over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_λ , roman_out end_POSTSUBSCRIPT end_ARG , (1)

where t𝑡titalic_t denotes time (or phase), indicating that high-resolution transmission spectra vary at different times due to Doppler shift and geometric effects (e.g., ingress/egress vs. full transit). To unpack each terms, we have

¯λ,outλ,s+¯λ,p(night),out(t),subscript¯𝜆outsubscript𝜆ssubscript¯𝜆pnightout𝑡\displaystyle\overline{\mathcal{F}}_{\mathrm{\lambda,out}}\equiv\mathcal{F}_{% \mathrm{\lambda,s}}+\overline{\mathcal{F}}_{\mathrm{\lambda,p(night),out}}(t),over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_λ , roman_out end_POSTSUBSCRIPT ≡ caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT + over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) , roman_out end_POSTSUBSCRIPT ( italic_t ) ,
λ,in(t)λ,sλ,s,obscured(t)+λ,p(night),in(t),subscript𝜆in𝑡subscript𝜆ssubscript𝜆sobscured𝑡subscript𝜆pnightin𝑡\displaystyle\mathcal{F}_{\mathrm{\lambda,in}}(t)\equiv\mathcal{F}_{\mathrm{% \lambda,s}}-\mathcal{F}_{\mathrm{\lambda,s,obscured}}(t)+\mathcal{F}_{\mathrm{% \lambda,p(night),in}}(t),caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_in end_POSTSUBSCRIPT ( italic_t ) ≡ caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT - caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_s , roman_obscured end_POSTSUBSCRIPT ( italic_t ) + caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) , roman_in end_POSTSUBSCRIPT ( italic_t ) , (2)

where “p(night)” denotes the planetary nightside, “s” denotes the star, “out” denotes out-of-transit, “in” denotes in-transit, and “obscured” denotes the stellar light obscured by planet (including atmosphere and disk). Substituting Equation (2.2) into Equation (1), we have

Δλ(t)=λ,s,obscured(t)+¯λ,p(night),out(t)λ,p(night),in(t)λ,s+¯λ,p(night),out(t).subscriptΔ𝜆𝑡subscript𝜆sobscured𝑡subscript¯𝜆pnightout𝑡subscript𝜆pnightin𝑡subscript𝜆ssubscript¯𝜆pnightout𝑡\Delta_{\mathrm{\lambda}}(t)=\frac{\mathcal{F}_{\mathrm{\lambda,s,obscured}}(t% )+\overline{\mathcal{F}}_{\mathrm{\lambda,p(night),out}}(t)-\mathcal{F}_{% \mathrm{\lambda,p(night),in}}(t)}{\mathcal{F}_{\mathrm{\lambda,s}}+\overline{% \mathcal{F}}_{\mathrm{\lambda,p(night),out}}(t)}.roman_Δ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_t ) = divide start_ARG caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_s , roman_obscured end_POSTSUBSCRIPT ( italic_t ) + over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) , roman_out end_POSTSUBSCRIPT ( italic_t ) - caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) , roman_in end_POSTSUBSCRIPT ( italic_t ) end_ARG start_ARG caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT + over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) , roman_out end_POSTSUBSCRIPT ( italic_t ) end_ARG . (3)

For high-resolution transmission spectroscopy, taking into account the Doppler effect from the variations in the line-of-sight velocity (see top panel of Figure 1), ¯λ,p(night),out(t)λ,p(night),in(t)subscript¯𝜆pnightout𝑡subscript𝜆pnightin𝑡\overline{\mathcal{F}}_{\mathrm{\lambda,p(night),out}}(t)\neq\mathcal{F}_{% \mathrm{\lambda,p(night),in}}(t)over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) , roman_out end_POSTSUBSCRIPT ( italic_t ) ≠ caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) , roman_in end_POSTSUBSCRIPT ( italic_t ).

Refer to caption
Figure 1: Top panel: Comparison between time-averaged nightside emission flux ¯λ,p(night),out(t)subscript¯𝜆pnightout𝑡\overline{\mathcal{F}}_{\mathrm{\lambda,p(night),out}}(t)over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) , roman_out end_POSTSUBSCRIPT ( italic_t ) and nightside emission flux at the moment of mid-transit λ,p(night),in(t=0)subscript𝜆pnightin𝑡0\mathcal{F}_{\mathrm{\lambda,p(night),in}}(t=0)caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) , roman_in end_POSTSUBSCRIPT ( italic_t = 0 ). Bottom panel: Comparison between classical and general transmission spectrum. The gray dashed line shows the classical transmission spectrum. The solid lines show our proposed general transmission spectra, which incorporate the nightside emission. The purple and cyan colors indicate that the nightside temperature structure is inverted and non-inverted, respectively. All models include rotational broadening.
Refer to caption
Figure 2: The H2O transmission spectrum model for WASP-33b. The upper left panel shows the model spectra of H2O, our general transmission spectrum (with nightside emission) in orange and the classical transmission spectrum (without nightside emission) in black. The lower left panel shows the normalized continuum-free model spectra. The middle panel shows the T-P profiles for the limb and nightside atmospheres. The gray shaded area shows the contribution function of the nightside thermal emission, integrated over the wavelength space. The dotted dashed lines show the condensation profiles of condensates commonly expected on high-temperature giants (Visscher et al., 2010; Wakeford et al., 2017). The peak of the contribution function here is above the intersection of the condensation profile with the nightside T-P, indicating that the thermal emission is not obscured by clouds. The right panel shows the H2O mass fraction calculated from the chemical equilibrium given by the T-P in the middle panel, compared to that of the nightside condensates.

We show in Appendix A the derivation of the general equation for high-resolution transmission spectrum with nighside emission. The one-dimensional (1D) general transmission spectrum (or called “transit depth”) can be expressed as,

Δλ(t)=[ελ,hetAλ,eff(t)AsLλ,p(night)(t)I¯λ,sAλ,eff(t)As]dλ,night,subscriptΔ𝜆𝑡delimited-[]subscript𝜀𝜆hetsubscript𝐴𝜆eff𝑡subscript𝐴ssubscript𝐿𝜆pnight𝑡subscript¯𝐼𝜆ssubscript𝐴𝜆eff𝑡subscript𝐴ssubscript𝑑𝜆night\displaystyle\Delta_{\lambda}(t)=\left[\varepsilon_{\mathrm{\lambda,het}}\frac% {A_{\mathrm{\lambda,eff}}(t)}{A_{\mathrm{s}}}-\frac{L_{\mathrm{\lambda,p(night% )}}(t)}{\bar{I}_{\mathrm{\lambda,s}}}\frac{A_{\mathrm{\lambda,eff}}(t)}{A_{% \mathrm{s}}}\right]\cdot d_{\mathrm{\lambda,night}},roman_Δ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_t ) = [ italic_ε start_POSTSUBSCRIPT italic_λ , roman_het end_POSTSUBSCRIPT divide start_ARG italic_A start_POSTSUBSCRIPT italic_λ , roman_eff end_POSTSUBSCRIPT ( italic_t ) end_ARG start_ARG italic_A start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT end_ARG - divide start_ARG italic_L start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT ( italic_t ) end_ARG start_ARG over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT end_ARG divide start_ARG italic_A start_POSTSUBSCRIPT italic_λ , roman_eff end_POSTSUBSCRIPT ( italic_t ) end_ARG start_ARG italic_A start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT end_ARG ] ⋅ italic_d start_POSTSUBSCRIPT italic_λ , roman_night end_POSTSUBSCRIPT , (4)

where dλ,night=1/[1+¯λ,p(night),out(t)/λ,s,out]=1/(1+FpRp2/FsRs2)subscript𝑑𝜆night1delimited-[]1subscript¯𝜆pnightout𝑡subscript𝜆sout11subscript𝐹psuperscriptsubscript𝑅p2subscript𝐹ssuperscriptsubscript𝑅s2d_{\mathrm{\lambda,night}}=1/[1+\overline{\mathcal{F}}_{\mathrm{\lambda,p(% night),out}}(t)/\mathcal{F}_{\mathrm{\lambda,s,out}}]=1/(1+F_{\mathrm{p}}R_{% \mathrm{p}}^{2}/F_{\mathrm{s}}R_{\mathrm{s}}^{2})italic_d start_POSTSUBSCRIPT italic_λ , roman_night end_POSTSUBSCRIPT = 1 / [ 1 + over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) , roman_out end_POSTSUBSCRIPT ( italic_t ) / caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_s , roman_out end_POSTSUBSCRIPT ] = 1 / ( 1 + italic_F start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_F start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) is the “nightside emission dilution factor” (Kipping & Tinetti, 2010). ελ,hetsubscript𝜀𝜆het\varepsilon_{\mathrm{\lambda,het}}italic_ε start_POSTSUBSCRIPT italic_λ , roman_het end_POSTSUBSCRIPT is the “unocculted stellar heterogeneities factor” (MacDonald & Lewis, 2022). Aλ,effsubscript𝐴𝜆effA_{\mathrm{\lambda,eff}}italic_A start_POSTSUBSCRIPT italic_λ , roman_eff end_POSTSUBSCRIPT and Assubscript𝐴sA_{\mathrm{s}}italic_A start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT denote the planetary effective area and stellar area, respectively. Lλ,p(night)=Iλ,p(night)I¯λ,p(night)subscript𝐿𝜆pnightsubscript𝐼𝜆pnightsubscript¯𝐼𝜆pnightL_{\mathrm{\lambda,p(night)}}=I_{\mathrm{\lambda,p(night)}}-\bar{I}_{\mathrm{% \lambda,p(night)}}italic_L start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT = italic_I start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT - over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT and I¯λ,ssubscript¯𝐼𝜆s\bar{I}_{\mathrm{\lambda,s}}over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT are the approximate spectral line intensity and mean stellar intensity, respectively. Lλ,p(night)(t)subscript𝐿𝜆pnight𝑡L_{\mathrm{\lambda,p(night)}}(t)italic_L start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT ( italic_t ) is the line intensity of the nightside emission at a given time after subtracting the continuum spectrum. For low-resolution transmission spectrum, Lλ,p(night)=0subscript𝐿𝜆pnight0L_{\mathrm{\lambda,p(night)}}=0italic_L start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT = 0, therefore, Equation (4) turns into Equation (1) of MacDonald & Lewis (2022).

In practice, assuming ελ,het=1subscript𝜀𝜆het1\varepsilon_{\mathrm{\lambda,het}}=1italic_ε start_POSTSUBSCRIPT italic_λ , roman_het end_POSTSUBSCRIPT = 1, we can rewrite Equation (4) as (omitting the notations t𝑡titalic_t and λ𝜆\mathrm{\lambda}italic_λ),

Tobssubscript𝑇obs\displaystyle T_{\mathrm{obs}}italic_T start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT =1Δabsent1Δ\displaystyle=1-\Delta= 1 - roman_Δ (5)
=1Reff2ψm(u)Rs2{1𝒢[Fpψe(u)]Fs}dλ,night,absent1superscriptsubscript𝑅eff2subscript𝜓m𝑢superscriptsubscript𝑅s21𝒢delimited-[]subscript𝐹psubscript𝜓e𝑢subscript𝐹ssubscript𝑑𝜆night\displaystyle=1-\frac{R_{\mathrm{eff}}^{2}\cdot\psi_{\mathrm{m}}(u)}{R_{% \mathrm{s}}^{2}}\left\{1-\frac{\mathcal{G}\left[F_{\mathrm{p}}\cdot\psi_{% \mathrm{e}}(u)\right]}{F_{\mathrm{s}}}\right\}\cdot d_{\mathrm{\lambda,night}},= 1 - divide start_ARG italic_R start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⋅ italic_ψ start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT ( italic_u ) end_ARG start_ARG italic_R start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG { 1 - divide start_ARG caligraphic_G [ italic_F start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT ⋅ italic_ψ start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT ( italic_u ) ] end_ARG start_ARG italic_F start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT end_ARG } ⋅ italic_d start_POSTSUBSCRIPT italic_λ , roman_night end_POSTSUBSCRIPT ,

where Tobssubscript𝑇obsT_{\mathrm{obs}}italic_T start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT denotes the general transmission spectrum model for subsequent analyses, and 𝒢𝒢\mathcal{G}caligraphic_G denotes Gaussian high-pass filter. Rssubscript𝑅sR_{\mathrm{s}}italic_R start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT and Fssubscript𝐹sF_{\mathrm{s}}italic_F start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT denote the stellar radius and flux (blackbody assumed), respectively. We use petitRADTRANS (Mollière et al., 2019) to implement the radiative transfer to compute Reffsubscript𝑅effR_{\mathrm{eff}}italic_R start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT and Fpsubscript𝐹pF_{\mathrm{p}}italic_F start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT. We set different two-point temperature structures for the Reffsubscript𝑅effR_{\mathrm{eff}}italic_R start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT and Fpsubscript𝐹pF_{\mathrm{p}}italic_F start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT calculations to account for the longitudinal variation in the temperature structure of the planetary atmosphere. The abundances of the chemical components are calculated from chemical equilibrium and solar elemental abundances under the assumed temperature structure, carbon-oxygen ratio (C/O), and metallicity ([Fe/H]). Our model also includes the continuum opacity of H- as well as the effects of condensation clouds. The bottom of Figure 1 presents an example that compares the general transmission spectrum with the classical transmission spectrum.

3 Observations and Methods

3.1 Observations and Data Reduction

We collected archived data observed by CARMENES (Quirrenbach et al., 2018) and GIANO-B (Claudi et al., 2017) for the transits of WASP-33b from the Calar Alto Archive111http://caha.sdc.cab.inta-csic.es/calto/jsp/searchform.jsp and the Italian Center for Astronomical Archive222http://archives.ia2.inaf.it/tng/, respectively. The CARMENES spectrograph is mounted on the 3.5 m telescope at Calar Alto Observatory. The CARMENES near-infrared channel has a wavelength coverage of 0.96–1.71 µmµm\micronroman_µm with high spectral resolution (R80,400similar-to𝑅80400R\sim 80,400italic_R ∼ 80 , 400). The GIARPS (Claudi et al., 2016) instrument, consisting of the GIANO-B spectrograph and the HARPS-North spectrograph, is installed on the Telescopio Nazionale Galileo, allowing simultaneous observation in the visible and near-infrared wavelengths. The GIANO-B spectrograph has a high resolution of R50,000similar-to𝑅50000R\sim 50,000italic_R ∼ 50 , 000 and covers the wavelength range from 0.95 to 2.45 µmµm\micronroman_µm. The basic method of high-resolution spectral data reduction follows Guilluy et al. (2019), Giacobbe et al. (2021), Yan et al. (2022a) and Yang et al. (2024). We used SYSREM (Tamuz et al., 2005) to remove telluric contamination and stellar spectral lines to obtain the final order-by-order spectral matrices for subsequent cross-correlation analyses (see Appendix B for the detailed data reduction process). The algorithm was introduced to analyze high-resolution spectroscopic time series by Birkby et al. (2013) and has become a standard techqniue in atmospheric characterization of (ultra-)hot Jupiters (e.g., Birkby et al., 2017; Nugroho et al., 2017; Hawker et al., 2018; Gibson et al., 2020; Yan et al., 2020).

3.2 Cross-Correlation

To search for the molecular spectral signal, we implemented the cross-correlation technique (Snellen et al., 2010). The transmission spectrum models (see Section 2.2) were convolved with a Gaussian function to match the instrumental resolution and resampled to the same wavelength grid as the observed spectra. The models were further normalized with a 12-point Gaussian high-pass filter to remove the broadband spectral features. The models were shifted from --150 to +++150 km s-1 with a step of 3 km s-1, which is larger than the pixel scale of both instruments to avoid oversampling. At each spectral shift, we computed the weighted cross-correlation function (CCF) by multiplying the weighted residual spectrum with the shifted spectral model. The CCF is formulated as:

CCF=rimi(\varv)σi2,CCFsubscript𝑟𝑖subscript𝑚𝑖\varvsuperscriptsubscript𝜎𝑖2\mathrm{CCF}=\sum\frac{r_{i}m_{i}(\varv)}{\sigma_{i}^{2}},roman_CCF = ∑ divide start_ARG italic_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( ) end_ARG start_ARG italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (6)

where risubscript𝑟𝑖r_{i}italic_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT represents the residual spectrum, mi(\varv)subscript𝑚𝑖\varvm_{i}(\varv)italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( ) is the spectral model shifted by velocity \varv\varv\varv, and σisubscript𝜎𝑖\sigma_{i}italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT denotes the error at the wavelength point i𝑖iitalic_i. For each residual spectrum, a corresponding CCF was derived. These individual CCFs were then combined by stacking to generate the CCF map for each observation and model.

To enhance the signal-to-noise ratio (S/N) of the planetary signals, the CCF map was converted from the terrestrial rest frame to the planetary rest frame using a grid of assumed planetary orbital velocity semi-amplitudes (Kpsubscript𝐾pK_{\mathrm{p}}italic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT). Assuming the planet has a circular orbit, the radial velocity (RV) of the planet \varvpsubscript\varvp\varv_{\mathrm{p}}start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT is expressed as

\varvp=\varvsys\varvbary+Kpsin(2πϕ)+Δ\varv,subscript\varvpsubscript\varvsyssubscript\varvbarysubscript𝐾p2𝜋italic-ϕΔ\varv\varv_{\mathrm{p}}=\varv_{\mathrm{sys}}-\varv_{\mathrm{bary}}+K_{\mathrm{p}}% \sin(2\pi\phi)+\Delta\varv,start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT = start_POSTSUBSCRIPT roman_sys end_POSTSUBSCRIPT - start_POSTSUBSCRIPT roman_bary end_POSTSUBSCRIPT + italic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT roman_sin ( 2 italic_π italic_ϕ ) + roman_Δ , (7)

where \varvsyssubscript\varvsys\varv_{\mathrm{sys}}start_POSTSUBSCRIPT roman_sys end_POSTSUBSCRIPT is the systemic velocity, \varvbarysubscript\varvbary\varv_{\mathrm{bary}}start_POSTSUBSCRIPT roman_bary end_POSTSUBSCRIPT is the barycentric Earth radial velocity (BERV), Δ\varvΔ\varv\Delta\varvroman_Δ is the radial velocity deviation from planetary rest frame, and ϕitalic-ϕ\phiitalic_ϕ is the orbital phase. A phase-folded one-dimensional CCF was generated for each assumed Kpsubscript𝐾pK_{\mathrm{p}}italic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT value by averaging the in-transit CCF map after being shifted to the planetary rest frame by that specific Kpsubscript𝐾pK_{\mathrm{p}}italic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT. A KpΔ\varvsubscript𝐾pΔ\varvK_{\mathrm{p}}-\Delta\varvitalic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT - roman_Δ map was then created by stacking the phase-folded one-dimensional CCFs corresponding to different Kpsubscript𝐾pK_{\mathrm{p}}italic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT values, ranging from 0 to 300 km s-1 with a step of 3 km s-1. The KpΔ\varvsubscript𝐾pΔ\varvK_{\mathrm{p}}-\Delta\varvitalic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT - roman_Δ map was normalized by the standard deviation calculated from +++0 to +++120 km s-1 in Kpsubscript𝐾pK_{\mathrm{p}}italic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT and +++30 to +++150 km s-1 in Δ\varvΔ\varv\Delta\varvroman_Δ. Consequently, the final KpΔ\varvsubscript𝐾pΔ\varvK_{\mathrm{p}}-\Delta\varvitalic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT - roman_Δ map was expressed in the form of S/N.

4 Results and Discussion

4.1 Evidence for H2O Emission During WASP-33b’s Transit

We employed the transmission spectrum models that incorporate planetary nightside thermal emission (as detailed in Section 2.2) to conduct a CCF analysis for WASP-33b using data from both CARMENES and GIANO-B (see Table 1 for the observational logs). We set up different T-P profiles at the limb and nightside to compute Reffsubscript𝑅effR_{\mathrm{eff}}italic_R start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT and Fpsubscript𝐹pF_{\mathrm{p}}italic_F start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT, respectively.

Refer to caption
Figure 3: CCF and KpΔ\varvsubscript𝐾pΔ\varvK_{\mathrm{p}}-\Delta\varvitalic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT - roman_Δ maps of H2O averaged over two nights of GIANO-B observations. First row: CCF map, where the two horizontal dashed lines indicate the start and end of the transit, and the tilted dashed line represents the expected planetary orbit. Second row: KpΔ\varvsubscript𝐾pΔ\varvK_{\mathrm{p}}-\Delta\varvitalic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT - roman_Δ map with orthogonal black dashed lines for the expected Kpsubscript𝐾pK_{\mathrm{p}}italic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT and Δ\varvΔ\varv\Delta\varvroman_Δ, and a pentagram for the position of the S/N peak. Third row: the one-dimensional CCF corresponding to the Kpsubscript𝐾pK_{\mathrm{p}}italic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT value at the S/N peak. The vertical gray dashed line indicates Δ\varv\text=0Δ\varv\text0\Delta\varv\text{=}0roman_Δ = 0.

We found evidence for the H2O emission features during the transit of WASP-33b. The transmission spectrum model we used is shown in Figure 2. The H2O emission signal is detected in the two GIANO-B transit observations with S/N \text=4.0\text4.0\text{=}4.0= 4.0 and S/N \text=5.5\text5.5\text{=}5.5= 5.5, respectively. By combining the observations from the two nights, we obtained the final KpΔ\varvsubscript𝐾pΔ\varvK_{\mathrm{p}}-\Delta\varvitalic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT - roman_Δ map shown in Figure 3. The H2O peak signal is located at Kp\text=22227+21\textkms1subscript𝐾p\textsuperscriptsubscript2222721\text𝑘𝑚superscript𝑠1K_{\mathrm{p}}\text{=}222_{-27}^{+21}\text{km~{}s}^{-1}italic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT = 222 start_POSTSUBSCRIPT - 27 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 21 end_POSTSUPERSCRIPT italic_k italic_m italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and Δ\varv\text=06+6\textkms1Δ\varv\textsuperscriptsubscript066\text𝑘𝑚superscript𝑠1\Delta\varv\text{=}0_{-6}^{+6}\text{km~{}s}^{-1}roman_Δ = 0 start_POSTSUBSCRIPT - 6 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 6 end_POSTSUPERSCRIPT italic_k italic_m italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT with an S/N of 5.3. This Kpsubscript𝐾pK_{\mathrm{p}}italic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT is consistent with the expected Kp\text=231\textkms1subscript𝐾p\text231\text𝑘𝑚superscript𝑠1K_{\mathrm{p}}\text{=}231\text{km~{}s}^{-1}italic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT = 231 italic_k italic_m italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT calculated from the orbital parameters of WASP-33b. A near-zero value for Δ\varvΔ\varv\Delta\varvroman_Δ suggests that there is no evidence for a strong day-night wind and/or an equatorial super-rotation jet in WASP-33b from the current observational data.

Previous optical transmission spectroscopy to detect atoms and ions in WASP-33b showed no significant circulation patterns (Yang et al., 2024; Yan et al., 2021), consistent with the results here. However, due to the strong influence of stellar pulsations, caution should be bared in mind when claiming species detection from the transmission spectrum of WASP-33b (Yang et al., 2024; Yan et al., 2021; Cauley et al., 2021; Borsa et al., 2021). Fortunately, H2O will not be present in WASP-33b’s host star due to the extremely high effective temperatures of A-type stars, so the effect of pulsations on the detection of H2O emission signals from WASP-33b should be negligible.

The observed S/N peaks of the H2O emission signal from the two nights are consistent at the 1.5σ𝜎\sigmaitalic_σ level (see Appendix C). The slight deviation could be due to insufficient data quality and imperfect data reduction, or could indicate the presence of planetary variability and a particular planetary climate pattern similar to WASP-121b (Changeat et al., 2024). On the other hand, we did not find robust emission signals from H2O in the CARMENES data, which we attribute to the insufficient coverage of the infrared band for CARMENES. We tested truncating the GIANO-B data at less-than-or-similar-to\lesssim1.75 µmµm\micronroman_µm to match the wavelength coverage of CARMENES and found that the detected emission signature would almost disappear (see Appendix D).

4.2 Does WASP-33b’s Nightside Have an Inverted Temperature Structure?

The inference of an emission signal indicates the existence of an inverted temperature structure on the nightside. According to the general understanding of hot Jupiters from theoretical models, the temperature difference between the dayside and the nightside essentially depends on the nature of the circulation, which is roughly determined by comparing the radiative timescale (τradsubscript𝜏rad\tau_{\mathrm{rad}}italic_τ start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT) with the wave propagation timescale or the advection timescale (τwavesubscript𝜏wave\tau_{\mathrm{wave}}italic_τ start_POSTSUBSCRIPT roman_wave end_POSTSUBSCRIPT or τadvsubscript𝜏adv\tau_{\mathrm{adv}}italic_τ start_POSTSUBSCRIPT roman_adv end_POSTSUBSCRIPT, Showman & Guillot, 2002; Komacek & Showman, 2016). In the equatorial region, there is no Coriolis force, with advection balancing the day-night pressure gradient force, and additional consideration of the Coriolis force is required at mid and high latitudes. As a result, as the planetary equilibrium temperature increases, τradsubscript𝜏rad\tau_{\mathrm{rad}}italic_τ start_POSTSUBSCRIPT roman_rad end_POSTSUBSCRIPT becomes shorter, leading to an increased day-night temperature difference. Such a pattern has already been confirmed by phase curve observations of hot Jupiters (Zhang, 2020).

However, compared to classical hot Jupiters, UHJs are subject to dissociation of molecules and ionization of atoms on the dayside due to higher dayside temperatures, and the radiative feedback from these processes affects the day-night heat transport, i.e. provides an additional energy budget. Several studies (Bell & Cowan, 2018; Komacek & Tan, 2018; Tan & Komacek, 2019) have suggested that the H2-H dissociation and recombination significantly increase the day-night heat transport, reducing the day-night temperature difference and the velocity of the equatorial super-rotation jets. In this process, H2 dissociates into H on the dayside, which cools the atmosphere; the H produced by the dissociation on the dayside is transported by the circulation to recombine into H2 on the nightside, which heats the atmosphere. If our H2O emission inference is correct, this new energy budget will make the nightside of the planet hotter than previously expected by the GCM model, indicating more efficient energy transport.

The presence of nightside thermal inversions seems to be an overheating of the upper atmosphere driven by atmospheric circulation. The mass mixing ratio of atomic hydrogen relative to the total gas, q=ρH/(ρH+ρH2)𝑞subscript𝜌Hsubscript𝜌Hsubscript𝜌subscriptH2q=\rho_{\mathrm{H}}/(\rho_{\mathrm{H}}+\rho_{\mathrm{H}_{2}})italic_q = italic_ρ start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT / ( italic_ρ start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT + italic_ρ start_POSTSUBSCRIPT roman_H start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ), increases with decreasing pressure, therefore corresponds to stronger horizontal heat transport at lower pressures. Meanwhile, the radiative cooling efficiency likely does not increase with decreasing pressure as fast as the heat transport. These two processes help to maintain a thermal inversion at part of the nightside. Due to the extremely short timescale of the H recombination process, chemical heating from this process occurs predominantly near the terminator of UHJs, as predicted by current state-of-the-art models.

Here, however, we find evidence for nightside thermal inversion, suggesting that more energy can be transported closer to the anti-stellar point. Tan et al. (2024) recently suggested that the formation of a nightside thermal inversion layer is possible when considering H2-H dissociation and recombination in planets with equilibrium temperatures above 2600 K and the absence of radiative cooling due to the lack of nightside coolants (e.g., TiO, VO). TiO has been detected on the dayside of WASP-33b by emission spectroscopy (Cont et al., 2021), but not on the terminator by transmission spectroscopy (Yang et al., 2024). This implies that the condensation of TiO occurs on the terminator/nightside and could then be circulated to the dayside to restore the gas phase. In addition, the heat transport and 3D thermal structure of UHJs are affected by other factors such as drag/damping processes (especially magnetic drag) and radiative feedback from nightside clouds.

An inverted temperature structure on the nightside is closely related to the nature of the atmospheric circulation, the efficiency of heat transport, and the specific energy budget. However, the 1D model ignores the effects of the 3D atmospheric structure and radiative transfer, as shown in Equation (A20) in Appendix A. The factor that dominates these effects is the rotation of the visible hemisphere during the observation. A portion of the limb and dayside atmosphere has already rotated into the visible hemisphere during the transit and is not fully visible on the nightside due to planetary rotation. We conducted a phase-resolved high-resolution simulation of H2O thermal emission spectra incorporating 3D velocity fields with PICASO (Batalha et al., 2019; Mukherjee et al., 2023) by post-processing the GCM outputs (see Appendix E). This simulation indicates that the 3D effects can cause the planetary thermal emission during transit to exhibit emission features for a planet as hot as WASP-33b (see Figure 9). As a result, the coupled transmission spectrum appears emission-like. This simulation supports our alternative speculation that the limb and dayside inverted temperature structures may also contribute to the emission features we detected.

In summary, we attribute the detected emission features to the nightside thermal inversion driven by atmospheric dynamics or the 3D effects. To fully interpret the detected emission features, it is necessary to consider the 3D atmospheric structure and radiative transfer effects to determine the precise boundary of the thermal inversion region. Further observations and the development of theories and models to better constrain the 3D chemistry, circulation, and thermal structure of UHJs are underway. WASP-33b is one of the best targets for testing these theories. Further quantitative investigations will be explored in a future paper.

Refer to caption
Figure 4: The general transmission spectrum models with three types of nightside T-P profiles for WASP-33b. The first three columns show the general transmission spectra of H2O, CO, and OH, respectively, while the fourth column shows the T-P profiles. Each inset shows a zoomed view of the spectrum. From top to bottom correspond to the inverted, non-inverted, and isothermal nightside T-P profiles.

4.3 Effects of Nightside Thermal Emission on Transmission Spectrum Observations

While the dilution effect of planetary nightside emission on transmission spectra has been discussed in several low-resolution observational studies (Stevenson et al., 2014; Kreidberg et al., 2018; Kirk et al., 2021), it has never been considered in the high-resolution transmission spectroscopy observations. We show that an approximation using the linear source functions at local thermodynamic equilibrium in Equations (A31)-(A33) in Appendix A. It follows that the effect of thermal emission on the transmission spectrum is determined by the nightside vertical temperature gradient and the ratio of the line and continuum absorption coefficients. That is, the larger the temperature gradient, line cross-section, and species abundance, the more significant the nightside thermal emission.

According to Section 2, we set up three types of temperature structures for the nightside of WASP-33b: (a) an inverted profile, (b) a non-inverted profile, and (c) an isothermal profile. For the limb, we simply assume an isothermal profile, although the limb temperature structure of the UHJ is generally inverted according to advanced GCM predictions, but the variation of the limb temperature structure is more a modulation of the atmospheric scale height and has a limited effect on the spectral line shape.

Figure 4 shows the transmission spectrum models with the three types of nightside temperature structures for H2O, CO and OH. We find that when considering the nightside thermal emission of WASP-33b, whenever the nightside is non-isothermal, it affects the transmission spectra of the molecules to different degrees. When the nightside has an inverted temperature structure, the H2O line shows pure emission features, while the other two molecules show absorption line wings and emission line cores, which weaken the transmission spectral signals, making the detection of these two molecules difficult. In contrast, the thermal emission produced by the non-inverted nightside temperature structure enhances the transmission spectral features of the three molecules.

In addition to these three molecules, the nightside thermal emission should also affect the transmission spectra of other molecules. It depends on the coincidence between the excitation temperatures of the spectral lines and the different temperature structures corresponding to different planets. We emphasize that the nightside thermal emission of UHJs should not be neglected in high-resolution infrared transit observations, and that at least some tests are required to assess the effects of the nightside thermal emission, especially in the case of non-isothermal nightside temperature structures.

Refer to caption
Figure 5: Comparison of KpΔ\varvsubscript𝐾pΔ\varvK_{\mathrm{p}}-\Delta\varvitalic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT - roman_Δ maps for the injection-recovery tests of WASP-33b. The injected general transmission model has included the nightside thermal emission. The cyan dashed lines indicate the expected values of the injected signal. The left and right panels show the results from the correct general transmission spectral template and the pure transmission spectral template, respectively. The first and second rows present the results for H2O and CO, respectively. The strength of the injected CO signal has been enhanced by five times.

Certainly, a non-negligible problem is the obscuring effect of clouds, but in the case of some hot Jupiters (UHJs or planets with high heat transport efficiency) the nightside temperatures are likely to be higher than the condensation temperatures of the various condensates, and the effect of the nightside thermal emission should not be neglected. Current GCM simulations show that the nightside cloud cover of UHJs is partial and dynamic, which is significantly different from the uniform cloud cover of hot Jupiters (Komacek et al., 2022). Moreover, the mean cloud top pressures on the nightside are similar-to\sim10–100 mbar, which is deeper than the temperature inversion caused by the thermal effects of hydrogen recombination at similar-to\sim1–10 mbar in GCMs. Therefore, the presence of these clouds would not affect the emission lines.

To illustrate how molecular detection would be affected if the nightside thermal emission of UHJs were ignored, we performed an injection-recovery test using H2O and CO as an example. The general transmission model created in the top panel of Figure 4 was injected into the observational data, and then a cross-correlation analysis was performed with a general transmission spectral template with nightside emission and a pure transmission spectral template without nightside emission, separately. The results are shown in Figure 5.

For H2O, a pure transmission spectral template neglecting the nightside thermal emission would lead to a non-detection. The H2O signal can only be recovered with the general transmission spectral template that includes the nightside thermal emission. For CO, the S/N of the current data cannot recover the injected signal even with the input general transmission model. It can be recovered only when the strength of the injected signal is increased five times. We note that neglecting the nightside thermal emission can lead to diametrically opposite features in the KpΔ\varvsubscript𝐾pΔ\varvK_{\mathrm{p}}-\Delta\varvitalic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT - roman_Δ map if the S/N is high enough or the signal strength is strong enough, which is due to the difference in the CO spectral line profiles between the general transmission model and the pure transmission model.

In summary, the nightside thermal emission could have a significant impact on the molecular detection of UHJs, and the nightside temperature structure and cloud shielding effect need to be further investigated. If clouds do not completely obscure the nightside thermal emission, the nightside thermal and chemical properties can be inferred directly from the transmission spectra using Equation (4).

5 Conclusions

We explore the influence of nightside thermal emission on high-resolution infrared transmission spectroscopy, especially in the case of UHJs. We present a general equation for the high-resolution transmission spectrum that includes planetary nightside thermal emission. This provides a new way to infer the thermal structure of the planetary nightside from high-resolution transmission spectroscopy observations. We find evidence for H2O emission in the UHJ WASP-33b during transit with the GIANO-B archival data, and interpret it as the presence of a nightside temperature inversion with a 1D general high-resolution transmission spectrum model. The S/N of the signal is not high enough to claim a robust detection, and further high-quality observations are needed to confirm our results.

We propose that some UHJs may exhibit non-isothermal temperature structures on the nightside, especially thermal inversions in the upper atmosphere due to higher than expected heat transport efficiencies, which would allow the nightside emission to penetrate the cloud layer. On the other hand, it is also possible that the 3D effect of the rotating visible hemisphere could cause a portion of the limb and dayside atmosphere to rotate into view during transit, resulting in emission features. Therefore, while it is necessary to consider the nightside thermal emission in high-resolution infrared transmission spectroscopy, a more in-depth study of the 3D chemical, circulation, and thermal structure of UHJs is also needed to fully understand the origin of a potential nightside thermal inversion.

The authors thank the anonymous reviewer for the constructive comments and suggestions on the manuscript. We thank Natasha Batalha for sharing the line-by-line opacity database used in computing the synthetic high-resolution spectrum in this work. G.C. acknowledges the support by the National Natural Science Foundation of China (NSFC; grant Nos. 12122308, 42075122), Youth Innovation Promotion Association CAS (2021315), and the Minor Planet Foundation of the Purple Mountain Observatory. F.Y. acknowledges the support by the NSFC (grant No. 42375118). J.J. acknowledges the support by the NSFC (grant No. 12033010).

References

  • Arcangeli et al. (2018) Arcangeli, J., Désert, J.-M., Line, M. R., et al. 2018, ApJL, 855, L30, doi: 10.3847/2041-8213/aab272
  • Batalha et al. (2019) Batalha, N. E., Marley, M. S., Lewis, N. K., & Fortney, J. J. 2019, ApJ, 878, 70, doi: 10.3847/1538-4357/ab1b51
  • Bell & Cowan (2018) Bell, T. J., & Cowan, N. B. 2018, ApJL, 857, L20, doi: 10.3847/2041-8213/aabcc8
  • Birkby et al. (2013) Birkby, J. L., de Kok, R. J., Brogi, M., et al. 2013, MNRAS, 436, L35, doi: 10.1093/mnrasl/slt107
  • Birkby et al. (2017) Birkby, J. L., de Kok, R. J., Brogi, M., Schwarz, H., & Snellen, I. A. G. 2017, AJ, 153, 138, doi: 10.3847/1538-3881/aa5c87
  • Borsa et al. (2021) Borsa, F., Lanza, A. F., Raspantini, I., et al. 2021, A&A, 653, A104, doi: 10.1051/0004-6361/202140559
  • Boucher et al. (2023) Boucher, A., Lafreniére, D., Pelletier, S., et al. 2023, MNRAS, 522, 5062, doi: 10.1093/mnras/stad1247
  • Brogi et al. (2018) Brogi, M., Giacobbe, P., Guilluy, G., et al. 2018, A&A, 615, A16, doi: 10.1051/0004-6361/201732189
  • Caballero et al. (2016) Caballero, J. A., Guàrdia, J., López del Fresno, M., et al. 2016, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 9910, Observatory Operations: Strategies, Processes, and Systems VI, ed. A. B. Peck, R. L. Seaman, & C. R. Benn, 99100E, doi: 10.1117/12.2233574
  • Carvalho & Johns-Krull (2023) Carvalho, A., & Johns-Krull, C. M. 2023, RNAAS, 7, 91, doi: 10.3847/2515-5172/acd37e
  • Cauley et al. (2021) Cauley, P. W., Wang, J., Shkolnik, E. L., et al. 2021, AJ, 161, 152, doi: 10.3847/1538-3881/abde43
  • Chakrabarty & Sengupta (2020) Chakrabarty, A., & Sengupta, S. 2020, ApJ, 898, 89, doi: 10.3847/1538-4357/ab9a33
  • Changeat et al. (2024) Changeat, Q., Skinner, J. W., Cho, J. Y. K., et al. 2024, ApJS, 270, 34, doi: 10.3847/1538-4365/ad1191
  • Chen et al. (2020) Chen, G., Casasayas-Barris, N., Pallé, E., et al. 2020, A&A, 635, A171, doi: 10.1051/0004-6361/201936986
  • Claudi et al. (2016) Claudi, R., Benatti, S., Carleo, I., et al. 2016, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 9908, Ground-based and Airborne Instrumentation for Astronomy VI, ed. C. J. Evans, L. Simard, & H. Takami, 99081A, doi: 10.1117/12.2231845
  • Claudi et al. (2017) Claudi, R., Benatti, S., Carleo, I., et al. 2017, EPJP, 132, 364, doi: 10.1140/epjp/i2017-11647-9
  • Collier Cameron et al. (2010) Collier Cameron, A., Guenther, E., Smalley, B., et al. 2010, MNRAS, 407, 507, doi: 10.1111/j.1365-2966.2010.16922.x
  • Cont et al. (2021) Cont, D., Yan, F., Reiners, A., et al. 2021, A&A, 651, A33, doi: 10.1051/0004-6361/202140732
  • Cowan & Agol (2011) Cowan, N. B., & Agol, E. 2011, ApJ, 726, 82, doi: 10.1088/0004-637X/726/2/82
  • Cowan et al. (2013) Cowan, N. B., Fuentes, P. A., & Haggard, H. M. 2013, MNRAS, 434, 2465, doi: 10.1093/mnras/stt1191
  • Fortney et al. (2008) Fortney, J. J., Lodders, K., Marley, M. S., & Freedman, R. S. 2008, ApJ, 678, 1419, doi: 10.1086/528370
  • Fortney et al. (2019) Fortney, J. J., Lupu, R. E., Morley, C. V., Freedman, R. S., & Hood, C. 2019, ApJL, 880, L16, doi: 10.3847/2041-8213/ab2a10
  • Fortney et al. (2005) Fortney, J. J., Marley, M. S., Lodders, K., Saumon, D., & Freedman, R. 2005, ApJL, 627, L69, doi: 10.1086/431952
  • Gandhi & Madhusudhan (2019) Gandhi, S., & Madhusudhan, N. 2019, MNRAS, 485, 5817, doi: 10.1093/mnras/stz751
  • Giacobbe et al. (2021) Giacobbe, P., Brogi, M., Gandhi, S., et al. 2021, Natur, 592, 205, doi: 10.1038/s41586-021-03381-x
  • Gibson et al. (2020) Gibson, N. P., Merritt, S., Nugroho, S. K., et al. 2020, MNRAS, 493, 2215, doi: 10.1093/mnras/staa228
  • Guilluy et al. (2019) Guilluy, G., Sozzetti, A., Brogi, M., et al. 2019, A&A, 625, A107, doi: 10.1051/0004-6361/201834615
  • Hawker et al. (2018) Hawker, G. A., Madhusudhan, N., Cabot, S. H. C., & Gandhi, S. 2018, ApJ, 863, L11, doi: 10.3847/2041-8213/aac49d
  • Hubeny et al. (2003) Hubeny, I., Burrows, A., & Sudarsky, D. 2003, ApJ, 594, 1011, doi: 10.1086/377080
  • Kipping & Tinetti (2010) Kipping, D. M., & Tinetti, G. 2010, MNRAS, 407, 2589, doi: 10.1111/j.1365-2966.2010.17094.x
  • Kirk et al. (2021) Kirk, J., Rackham, B. V., MacDonald, R. J., et al. 2021, AJ, 162, 34, doi: 10.3847/1538-3881/abfcd2
  • Komacek & Showman (2016) Komacek, T. D., & Showman, A. P. 2016, ApJ, 821, 16, doi: 10.3847/0004-637X/821/1/16
  • Komacek & Tan (2018) Komacek, T. D., & Tan, X. 2018, RNAAS, 2, 36, doi: 10.3847/2515-5172/aac5e7
  • Komacek et al. (2022) Komacek, T. D., Tan, X., Gao, P., & Lee, E. K. H. 2022, ApJ, 934, 79, doi: 10.3847/1538-4357/ac7723
  • Kreidberg et al. (2018) Kreidberg, L., Line, M. R., Parmentier, V., et al. 2018, AJ, 156, 17, doi: 10.3847/1538-3881/aac3df
  • Langeveld et al. (2021) Langeveld, A. B., Madhusudhan, N., Cabot, S. H. C., & Hodgkin, S. T. 2021, MNRAS, 502, 4392, doi: 10.1093/mnras/stab134
  • Lothringer et al. (2018) Lothringer, J. D., Barman, T., & Koskinen, T. 2018, ApJ, 866, 27, doi: 10.3847/1538-4357/aadd9e
  • MacDonald & Lewis (2022) MacDonald, R. J., & Lewis, N. K. 2022, ApJ, 929, 20, doi: 10.3847/1538-4357/ac47fe
  • Martin-Lagarde et al. (2020) Martin-Lagarde, M., Morello, G., Lagage, P.-O., Gastaud, R., & Cossou, C. 2020, AJ, 160, 197, doi: 10.3847/1538-3881/abac09
  • Mollière et al. (2019) Mollière, P., Wardenier, J. P., van Boekel, R., et al. 2019, A&A, 627, A67, doi: 10.1051/0004-6361/201935470
  • Morello et al. (2021) Morello, G., Zingales, T., Martin-Lagarde, M., Gastaud, R., & Lagage, P.-O. 2021, AJ, 161, 174, doi: 10.3847/1538-3881/abe048
  • Mukherjee et al. (2023) Mukherjee, S., Batalha, N. E., Fortney, J. J., & Marley, M. S. 2023, ApJ, 942, 71, doi: 10.3847/1538-4357/ac9f48
  • Nugroho et al. (2017) Nugroho, S. K., Kawahara, H., Masuda, K., et al. 2017, AJ, 154, 221, doi: 10.3847/1538-3881/aa9433
  • Parmentier et al. (2015) Parmentier, V., Guillot, T., Fortney, J. J., & Marley, M. S. 2015, A&A, 574, A35, doi: 10.1051/0004-6361/201323127
  • Parmentier et al. (2018) Parmentier, V., Line, M. R., Bean, J. L., et al. 2018, A&A, 617, A110, doi: 10.1051/0004-6361/201833059
  • Quirrenbach et al. (2018) Quirrenbach, A., Amado, P. J., Ribas, I., et al. 2018, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 10702, Ground-based and Airborne Instrumentation for Astronomy VII, ed. C. J. Evans, L. Simard, & H. Takami, 107020W, doi: 10.1117/12.2313689
  • Rainer et al. (2018) Rainer, M., Harutyunyan, A., Carleo, I., et al. 2018, in Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol. 10702, Ground-based and Airborne Instrumentation for Astronomy VII, ed. C. J. Evans, L. Simard, & H. Takami, 1070266, doi: 10.1117/12.2312130
  • Rauscher et al. (2018) Rauscher, E., Suri, V., & Cowan, N. B. 2018, AJ, 156, 235, doi: 10.3847/1538-3881/aae57f
  • Showman & Guillot (2002) Showman, A. P., & Guillot, T. 2002, A&A, 385, 166, doi: 10.1051/0004-6361:20020101
  • Snellen et al. (2010) Snellen, I. A. G., de Kok, R. J., de Mooij, E. J. W., & Albrecht, S. 2010, Natur, 465, 1049, doi: 10.1038/nature09111
  • Stevenson et al. (2014) Stevenson, K. B., Bean, J. L., Seifahrt, A., et al. 2014, AJ, 147, 161, doi: 10.1088/0004-6256/147/6/161
  • Tamuz et al. (2005) Tamuz, O., Mazeh, T., & Zucker, S. 2005, MNRAS, 356, 1466, doi: 10.1111/j.1365-2966.2004.08585.x
  • Tan & Komacek (2019) Tan, X., & Komacek, T. D. 2019, ApJ, 886, 26, doi: 10.3847/1538-4357/ab4a76
  • Tan et al. (2024) Tan, X., Komacek, T. D., Batalha, N. E., et al. 2024, MNRAS, 528, 1016, doi: 10.1093/mnras/stae050
  • Toon et al. (1989) Toon, O. B., McKay, C. P., Ackerman, T. P., & Santhanam, K. 1989, J. Geophys. Res., 94, 16287, doi: 10.1029/JD094iD13p16287
  • Visscher et al. (2010) Visscher, C., Lodders, K., & Fegley, Bruce, J. 2010, ApJ, 716, 1060, doi: 10.1088/0004-637X/716/2/1060
  • Wakeford et al. (2017) Wakeford, H. R., Visscher, C., Lewis, N. K., et al. 2017, MNRAS, 464, 4247, doi: 10.1093/mnras/stw2639
  • Yan et al. (2020) Yan, F., Pallé, E., Reiners, A., et al. 2020, A&A, 640, L5, doi: 10.1051/0004-6361/202038294
  • Yan et al. (2021) Yan, F., Wyttenbach, A., Casasayas-Barris, N., et al. 2021, A&A, 645, A22, doi: 10.1051/0004-6361/202039302
  • Yan et al. (2022a) Yan, F., Reiners, A., Pallé, E., et al. 2022a, A&A, 659, A7, doi: 10.1051/0004-6361/202142395
  • Yan et al. (2022b) Yan, F., Pallé, E., Reiners, A., et al. 2022b, A&A, 661, L6, doi: 10.1051/0004-6361/202243503
  • Yang et al. (2024) Yang, Y., Chen, G., Wang, S., & Yan, F. 2024, AJ, 167, 36, doi: 10.3847/1538-3881/ad10a3
  • Zhang et al. (2018) Zhang, M., Knutson, H. A., Kataria, T., et al. 2018, AJ, 155, 83, doi: 10.3847/1538-3881/aaa458
  • Zhang (2020) Zhang, X. 2020, RAA, 20, 099, doi: 10.1088/1674-4527/20/7/99

Appendix A Derivation of a Time-Resolved General Equation for Exoplanet High-Resolution Transmission Spectra.

Here we derive the general equation for high-resolution transmission spectrum that incorporates planetary nightside thermal emission shown in Equation (4) and simplify the equation to illustrate the effect of planetary nightside emission on the transmission spectrum. First we derive a formal equation for the time-resolved high-resolution transmission spectra for 3D atmospheric and radiative transfer calculations. We then focus on the case of the 1D atmospheric solution and present an approximate solution for a special situation. This derivation builds on the concept of MacDonald & Lewis (2022) for the low-resolution transmission spectrum.

The transmission spectrum is defined as the spectral flux difference between the out-of-transit time and the in-transit time. At high resolution, the transmission spectrum will be Doppler-shifted at each time (or phase) due to the orbital motion of the planet. Therefore, the formal equation for high-resolution transmission spectra will be a time-dependent equation:

Δλ(t)¯λ,out(t)λ,in(t)¯λ,out(t).subscriptΔ𝜆𝑡subscript¯𝜆out𝑡subscript𝜆in𝑡subscript¯𝜆out𝑡\Delta_{\lambda}(t)\equiv\frac{\overline{\mathcal{F}}_{\mathrm{\lambda,out}}(t% )-\mathcal{F}_{\mathrm{\lambda,in}}(t)}{\overline{\mathcal{F}}_{\mathrm{% \lambda,out}}(t)}.roman_Δ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_t ) ≡ divide start_ARG over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_λ , roman_out end_POSTSUBSCRIPT ( italic_t ) - caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_in end_POSTSUBSCRIPT ( italic_t ) end_ARG start_ARG over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_λ , roman_out end_POSTSUBSCRIPT ( italic_t ) end_ARG . (A1)

To unpack each terms, we have

λ,out(t)λ,s,out(t)+λ,p(night),out(t),subscript𝜆out𝑡subscript𝜆sout𝑡subscript𝜆pnightout𝑡\displaystyle\mathcal{F}_{\mathrm{\lambda,out}}(t)\equiv\mathcal{F}_{\mathrm{% \lambda,s,out}}(t)+\mathcal{F}_{\mathrm{\lambda,p(night),out}}(t),caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_out end_POSTSUBSCRIPT ( italic_t ) ≡ caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_s , roman_out end_POSTSUBSCRIPT ( italic_t ) + caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) , roman_out end_POSTSUBSCRIPT ( italic_t ) ,
λ,in(t)λ,s,in(t)+λ,s,trans(t)+λ,p(night),in(t),subscript𝜆in𝑡subscript𝜆sin𝑡subscript𝜆strans𝑡subscript𝜆pnightin𝑡\displaystyle\mathcal{F}_{\mathrm{\lambda,in}}(t)\equiv\mathcal{F}_{\mathrm{% \lambda,s,in}}(t)+\mathcal{F}_{\mathrm{\lambda,s,trans}}(t)+\mathcal{F}_{% \mathrm{\lambda,p(night),in}}(t),caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_in end_POSTSUBSCRIPT ( italic_t ) ≡ caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_s , roman_in end_POSTSUBSCRIPT ( italic_t ) + caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_s , roman_trans end_POSTSUBSCRIPT ( italic_t ) + caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) , roman_in end_POSTSUBSCRIPT ( italic_t ) , (A2)

where “p(night)” denotes the planetary nightside, “s” denotes the star, “out” denotes out-of-transit, “in” denotes in-transit, and “trans” denotes the transmission of stellar light through the planetary atmosphere. The average flux outside of transit can be expressed as

¯λ,out(t)=λ,s,out+¯λ,p(night),out(t),subscript¯𝜆out𝑡subscript𝜆soutsubscript¯𝜆pnightout𝑡\overline{\mathcal{F}}_{\mathrm{\lambda,out}}(t)=\mathcal{F}_{\mathrm{\lambda,% s,out}}+\overline{\mathcal{F}}_{\mathrm{\lambda,p(night),out}}(t),over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_λ , roman_out end_POSTSUBSCRIPT ( italic_t ) = caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_s , roman_out end_POSTSUBSCRIPT + over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) , roman_out end_POSTSUBSCRIPT ( italic_t ) , (A3)

where ¯λ,p(night),out(t)subscript¯𝜆pnightout𝑡\overline{\mathcal{F}}_{\mathrm{\lambda,p(night),out}}(t)over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) , roman_out end_POSTSUBSCRIPT ( italic_t ) is time-averaged planetary nightside thermal flux outside of transit. Substituting Equation (A) and (A3) into Equation (A1), we have

Δλ(t)=λ,s,outλ,s,in(t)λ,s,trans(t)+¯λ,p(night),out(t)λ,p(night),in(t)λ,s,out+¯λ,p(night),out(t).subscriptΔ𝜆𝑡subscript𝜆soutsubscript𝜆sin𝑡subscript𝜆strans𝑡subscript¯𝜆pnightout𝑡subscript𝜆pnightin𝑡subscript𝜆soutsubscript¯𝜆pnightout𝑡\Delta_{\mathrm{\lambda}}(t)=\frac{\mathcal{F}_{\mathrm{\lambda,s,out}}-% \mathcal{F}_{\mathrm{\lambda,s,in}}(t)-\mathcal{F}_{\mathrm{\lambda,s,trans}}(% t)+\overline{\mathcal{F}}_{\mathrm{\lambda,p(night),out}}(t)-\mathcal{F}_{% \mathrm{\lambda,p(night),in}}(t)}{\mathcal{F}_{\mathrm{\lambda,s,out}}+% \overline{\mathcal{F}}_{\mathrm{\lambda,p(night),out}}(t)}.roman_Δ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_t ) = divide start_ARG caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_s , roman_out end_POSTSUBSCRIPT - caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_s , roman_in end_POSTSUBSCRIPT ( italic_t ) - caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_s , roman_trans end_POSTSUBSCRIPT ( italic_t ) + over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) , roman_out end_POSTSUBSCRIPT ( italic_t ) - caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) , roman_in end_POSTSUBSCRIPT ( italic_t ) end_ARG start_ARG caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_s , roman_out end_POSTSUBSCRIPT + over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) , roman_out end_POSTSUBSCRIPT ( italic_t ) end_ARG . (A4)

The flux is defined as,

λ=ΩIλ𝐧^𝐤^𝑑Ω,subscript𝜆subscriptΩsubscript𝐼𝜆^𝐧^𝐤differential-dΩ\mathcal{F}_{\lambda}=\int_{\Omega}I_{\lambda}\hat{\mathbf{n}}\cdot\hat{% \mathbf{k}}d\Omega,caligraphic_F start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT = ∫ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT italic_I start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT over^ start_ARG bold_n end_ARG ⋅ over^ start_ARG bold_k end_ARG italic_d roman_Ω , (A5)

where Iλsubscript𝐼𝜆I_{\mathrm{\lambda}}italic_I start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT is the spectral intensity, ΩΩ\Omegaroman_Ω is the solid angle subtended by the source at the observer, and 𝐧^^𝐧\hat{\mathbf{n}}over^ start_ARG bold_n end_ARG and 𝐤^^𝐤\hat{\mathbf{k}}over^ start_ARG bold_k end_ARG are the unit vectors in the direction of the normal and observer, respectively. The stellar rays satisfy 𝐧^𝐤^=1^𝐧^𝐤1\hat{\mathbf{n}}\cdot\hat{\mathbf{k}}=1over^ start_ARG bold_n end_ARG ⋅ over^ start_ARG bold_k end_ARG = 1 in the transit geometry, but the planetary thermal emission does not. Therefore, we have

λ,s,outsubscript𝜆sout\displaystyle\mathcal{F}_{\mathrm{\lambda,s,out}}caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_s , roman_out end_POSTSUBSCRIPT =Ωs(full)Iλ,s𝑑Ω,absentsubscriptsubscriptΩsfullsubscript𝐼𝜆sdifferential-dΩ\displaystyle=\int_{\Omega_{\mathrm{s(full)}}}I_{\mathrm{\lambda,s}}d\Omega,= ∫ start_POSTSUBSCRIPT roman_Ω start_POSTSUBSCRIPT roman_s ( roman_full ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_I start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT italic_d roman_Ω , (A6)
λ,s,in(t)subscript𝜆sin𝑡\displaystyle\mathcal{F}_{\mathrm{\lambda,s,in}}(t)caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_s , roman_in end_POSTSUBSCRIPT ( italic_t ) =Ωs,unobscured(t)Iλ,s𝑑Ω,absentsubscriptsubscriptΩsunobscured𝑡subscript𝐼𝜆sdifferential-dΩ\displaystyle=\int_{\Omega_{\mathrm{s,unobscured}}(t)}I_{\mathrm{\lambda,s}}d\Omega,= ∫ start_POSTSUBSCRIPT roman_Ω start_POSTSUBSCRIPT roman_s , roman_unobscured end_POSTSUBSCRIPT ( italic_t ) end_POSTSUBSCRIPT italic_I start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT italic_d roman_Ω , (A7)
λ,s,trans(t)subscript𝜆strans𝑡\displaystyle\mathcal{F}_{\mathrm{\lambda,s,trans}}(t)caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_s , roman_trans end_POSTSUBSCRIPT ( italic_t ) =Ωp(t)Iλ,s,trans(t)𝑑Ω,absentsubscriptsubscriptΩp𝑡subscript𝐼𝜆strans𝑡differential-dΩ\displaystyle=\int_{\Omega_{\mathrm{p}}(t)}I_{\mathrm{\lambda,s,trans}}(t)d\Omega,= ∫ start_POSTSUBSCRIPT roman_Ω start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT ( italic_t ) end_POSTSUBSCRIPT italic_I start_POSTSUBSCRIPT italic_λ , roman_s , roman_trans end_POSTSUBSCRIPT ( italic_t ) italic_d roman_Ω , (A8)
λ,p(night)(t)subscript𝜆pnight𝑡\displaystyle\mathcal{F}_{\mathrm{\lambda,p(night)}}(t)caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT ( italic_t ) =V(θ,ϕ,t)Iλ,p(θ,ϕ,t)𝑑Ω,absentcontour-integral𝑉𝜃italic-ϕ𝑡subscript𝐼𝜆p𝜃italic-ϕ𝑡differential-dΩ\displaystyle=\oint V(\theta,\phi,t)I_{\mathrm{\lambda,p}}(\theta,\phi,t)d\Omega,= ∮ italic_V ( italic_θ , italic_ϕ , italic_t ) italic_I start_POSTSUBSCRIPT italic_λ , roman_p end_POSTSUBSCRIPT ( italic_θ , italic_ϕ , italic_t ) italic_d roman_Ω , (A9)

where the visibility function (see Rauscher et al., 2018; Cowan et al., 2013; Cowan & Agol, 2011) is

V(θ,ϕ,t)=max[cosγ0,0]=max[sinθsinθ0cos(ϕϕ0)+cosθcosθ0,0],𝑉𝜃italic-ϕ𝑡subscript𝛾00𝜃subscript𝜃0italic-ϕsubscriptitalic-ϕ0𝜃subscript𝜃00V(\theta,\phi,t)=\max\left[\cos\gamma_{0},0\right]=\max\left[\sin\theta\sin% \theta_{0}\cos\left(\phi-\phi_{0}\right)+\cos\theta\cos\theta_{0},0\right],italic_V ( italic_θ , italic_ϕ , italic_t ) = roman_max [ roman_cos italic_γ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , 0 ] = roman_max [ roman_sin italic_θ roman_sin italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_cos ( italic_ϕ - italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) + roman_cos italic_θ roman_cos italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , 0 ] , (A10)

where θ𝜃\thetaitalic_θ and ϕitalic-ϕ\phiitalic_ϕ are the latitude and longitude, respectively, θ0(t)=θ0subscript𝜃0𝑡subscript𝜃0\theta_{0}(t)=\theta_{0}italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_t ) = italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, ϕ0(t)=ϕ0(0)ωrottsubscriptitalic-ϕ0𝑡subscriptitalic-ϕ00subscript𝜔rot𝑡\phi_{0}(t)=\phi_{0}(0)-\omega_{\mathrm{rot}}titalic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_t ) = italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 0 ) - italic_ω start_POSTSUBSCRIPT roman_rot end_POSTSUBSCRIPT italic_t, and ωrotsubscript𝜔rot\omega_{\mathrm{rot}}italic_ω start_POSTSUBSCRIPT roman_rot end_POSTSUBSCRIPT is planetary rotation angular velocity. For tidally locked planets, ωrot=ωorbitsubscript𝜔rotsubscript𝜔orbit\omega_{\mathrm{rot}}=\omega_{\mathrm{orbit}}italic_ω start_POSTSUBSCRIPT roman_rot end_POSTSUBSCRIPT = italic_ω start_POSTSUBSCRIPT roman_orbit end_POSTSUBSCRIPT. Moreover, we can rewrite

λ,s,outλ,s,in(t)subscript𝜆soutsubscript𝜆sin𝑡\displaystyle\mathcal{F}_{\mathrm{\lambda,s,out}}-\mathcal{F}_{\mathrm{\lambda% ,s,in}}(t)caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_s , roman_out end_POSTSUBSCRIPT - caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_s , roman_in end_POSTSUBSCRIPT ( italic_t ) =Ωoverlap(t)Iλ,s𝑑Ω.absentsubscriptsubscriptΩoverlap𝑡subscript𝐼𝜆sdifferential-dΩ\displaystyle=\int_{\Omega_{\mathrm{overlap}}(t)}I_{\mathrm{\lambda,s}}d\Omega.= ∫ start_POSTSUBSCRIPT roman_Ω start_POSTSUBSCRIPT roman_overlap end_POSTSUBSCRIPT ( italic_t ) end_POSTSUBSCRIPT italic_I start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT italic_d roman_Ω . (A11)

Then, we introduce the mean “stellar intensity” and the time-averaged “planetary thermal emission intensity”:

I¯λ,ssubscript¯𝐼𝜆s\displaystyle\overline{I}_{\mathrm{\lambda,s}}over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT =Ωs(full)Iλ,s𝑑ΩΩs(full)𝑑Ω,absentsubscriptsubscriptΩsfullsubscript𝐼𝜆sdifferential-dΩsubscriptsubscriptΩsfulldifferential-dΩ\displaystyle=\frac{\int_{\Omega_{\mathrm{s(full)}}}I_{\mathrm{\lambda,s}}d% \Omega}{\int_{\Omega_{\mathrm{s(full)}}}d\Omega},= divide start_ARG ∫ start_POSTSUBSCRIPT roman_Ω start_POSTSUBSCRIPT roman_s ( roman_full ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_I start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT italic_d roman_Ω end_ARG start_ARG ∫ start_POSTSUBSCRIPT roman_Ω start_POSTSUBSCRIPT roman_s ( roman_full ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_d roman_Ω end_ARG , (A12)
I¯λ,p(night)subscript¯𝐼𝜆pnight\displaystyle\overline{I}_{\mathrm{\lambda,p(night)}}over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT =V(θ,ϕ,t)Iλ,p(night)(θ,ϕ,t)𝑑Ω𝑑tV(θ,ϕ,t)𝑑Ω𝑑t.absentcontour-integral𝑉𝜃italic-ϕ𝑡subscript𝐼𝜆pnight𝜃italic-ϕ𝑡differential-dΩdifferential-d𝑡contour-integral𝑉𝜃italic-ϕ𝑡differential-dΩdifferential-d𝑡\displaystyle=\frac{\int\oint V(\theta,\phi,t)I_{\mathrm{\lambda,p(night)}}(% \theta,\phi,t)d\Omega dt}{\int\oint V(\theta,\phi,t)d\Omega dt}.= divide start_ARG ∫ ∮ italic_V ( italic_θ , italic_ϕ , italic_t ) italic_I start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT ( italic_θ , italic_ϕ , italic_t ) italic_d roman_Ω italic_d italic_t end_ARG start_ARG ∫ ∮ italic_V ( italic_θ , italic_ϕ , italic_t ) italic_d roman_Ω italic_d italic_t end_ARG .

Therefore,

λ,s,outsubscript𝜆sout\displaystyle\mathcal{F}_{\mathrm{\lambda,s,out}}caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_s , roman_out end_POSTSUBSCRIPT =I¯λ,sΩs(full),absentsubscript¯𝐼𝜆ssubscriptΩsfull\displaystyle=\bar{I}_{\mathrm{\lambda,s}}\Omega_{\mathrm{s(full)}},= over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT roman_Ω start_POSTSUBSCRIPT roman_s ( roman_full ) end_POSTSUBSCRIPT , (A13)
¯λ,p(night),out(t)subscript¯𝜆pnightout𝑡\displaystyle\overline{\mathcal{F}}_{\mathrm{\lambda,p(night),out}}(t)over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) , roman_out end_POSTSUBSCRIPT ( italic_t ) =I¯λ,p(night)Ωp(full).absentsubscript¯𝐼𝜆pnightsubscriptΩpfull\displaystyle=\overline{I}_{\mathrm{\lambda,p(night)}}\Omega_{\mathrm{p(full)}}.= over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT roman_Ω start_POSTSUBSCRIPT roman_p ( roman_full ) end_POSTSUBSCRIPT .

Next, we replace the solid angle integral with the projected area integral. Ω(full)=A/D2subscriptΩfull𝐴superscript𝐷2\Omega_{\mathrm{(full)}}=A/D^{2}roman_Ω start_POSTSUBSCRIPT ( roman_full ) end_POSTSUBSCRIPT = italic_A / italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, so dΩ(full)=dA/D2𝑑subscriptΩfull𝑑𝐴superscript𝐷2d\Omega_{\mathrm{(full)}}=dA/D^{2}italic_d roman_Ω start_POSTSUBSCRIPT ( roman_full ) end_POSTSUBSCRIPT = italic_d italic_A / italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, where A=πR2𝐴𝜋superscript𝑅2A=\pi R^{2}italic_A = italic_π italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is the projected area and D𝐷Ditalic_D is the distance from the star to the observer. Therefore, substituting Equation (A6)-(A13) into Equation (A4), we have

Δλ(t)=subscriptΔ𝜆𝑡absent\displaystyle\Delta_{\lambda}(t)=roman_Δ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_t ) = [Aoverlap(t)Iλ,sI¯λ,s𝑑AAp(full)Iλ,s,trans(t)I¯λ,s𝑑A+Ap(full)I¯λ,p(night)I¯λ,sAp(full)V(θ,ϕ,t)Iλ,p(night)(θ,ϕ,t)I¯λ,s𝑑A]delimited-[]subscriptsubscript𝐴overlap𝑡subscript𝐼𝜆𝑠subscript¯𝐼𝜆𝑠differential-d𝐴subscriptsubscript𝐴pfullsubscript𝐼𝜆strans𝑡subscript¯𝐼𝜆𝑠differential-d𝐴subscript𝐴pfullsubscript¯𝐼𝜆pnightsubscript¯𝐼𝜆ssubscriptsubscript𝐴pfull𝑉𝜃italic-ϕ𝑡subscript𝐼𝜆pnight𝜃italic-ϕ𝑡subscript¯𝐼𝜆𝑠differential-d𝐴\displaystyle\left[\int_{A_{\mathrm{overlap}}(t)}\frac{I_{\lambda,s}}{% \overline{I}_{\lambda,s}}dA-\int_{A_{\mathrm{p(full)}}}\frac{I_{\mathrm{% \lambda,s,trans}}(t)}{\overline{I}_{\lambda,s}}dA+A_{\mathrm{p(full)}}\frac{% \overline{I}_{\mathrm{\lambda,p(night)}}}{\overline{I}_{\mathrm{\lambda,s}}}-% \int_{A_{\mathrm{p(full)}}}\frac{V(\theta,\phi,t)I_{\mathrm{\lambda,p(night)}}% (\theta,\phi,t)}{\overline{I}_{\lambda,s}}dA\right][ ∫ start_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT roman_overlap end_POSTSUBSCRIPT ( italic_t ) end_POSTSUBSCRIPT divide start_ARG italic_I start_POSTSUBSCRIPT italic_λ , italic_s end_POSTSUBSCRIPT end_ARG start_ARG over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , italic_s end_POSTSUBSCRIPT end_ARG italic_d italic_A - ∫ start_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT roman_p ( roman_full ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT divide start_ARG italic_I start_POSTSUBSCRIPT italic_λ , roman_s , roman_trans end_POSTSUBSCRIPT ( italic_t ) end_ARG start_ARG over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , italic_s end_POSTSUBSCRIPT end_ARG italic_d italic_A + italic_A start_POSTSUBSCRIPT roman_p ( roman_full ) end_POSTSUBSCRIPT divide start_ARG over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT end_ARG start_ARG over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT end_ARG - ∫ start_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT roman_p ( roman_full ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT divide start_ARG italic_V ( italic_θ , italic_ϕ , italic_t ) italic_I start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT ( italic_θ , italic_ϕ , italic_t ) end_ARG start_ARG over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , italic_s end_POSTSUBSCRIPT end_ARG italic_d italic_A ] (A14)
×1As(full)+Ap(full)I¯λ,p(night)I¯λ,s.absent1subscript𝐴sfullsubscript𝐴pfullsubscript¯𝐼𝜆pnightsubscript¯𝐼𝜆s\displaystyle\times\frac{1}{A_{\mathrm{s(full)}}+A_{\mathrm{p(full)}}\frac{% \overline{I}_{\mathrm{\lambda,p(night)}}}{\overline{I}_{\mathrm{\lambda,s}}}}.× divide start_ARG 1 end_ARG start_ARG italic_A start_POSTSUBSCRIPT roman_s ( roman_full ) end_POSTSUBSCRIPT + italic_A start_POSTSUBSCRIPT roman_p ( roman_full ) end_POSTSUBSCRIPT divide start_ARG over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT end_ARG start_ARG over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT end_ARG end_ARG .

Next, we follow MacDonald & Lewis (2022) to reduce Iλ,s,transsubscript𝐼𝜆stransI_{\mathrm{\lambda,s,trans}}italic_I start_POSTSUBSCRIPT italic_λ , roman_s , roman_trans end_POSTSUBSCRIPT, we have

Iλ,s,trans(t)subscript𝐼𝜆strans𝑡\displaystyle I_{\mathrm{\lambda,s,trans}}(t)italic_I start_POSTSUBSCRIPT italic_λ , roman_s , roman_trans end_POSTSUBSCRIPT ( italic_t ) =m=1Nphotδray,s,m𝒯λ,m(t)Iλ,s,mabsentsuperscriptsubscript𝑚1subscript𝑁photsubscript𝛿raysmsubscript𝒯𝜆m𝑡subscript𝐼𝜆sm\displaystyle=\sum_{m=1}^{N_{\mathrm{phot}}}\delta_{\mathrm{ray,s,m}}\mathcal{% T}_{\mathrm{\lambda,m}}(t)I_{\mathrm{\lambda,s,m}}= ∑ start_POSTSUBSCRIPT italic_m = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT roman_phot end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_δ start_POSTSUBSCRIPT roman_ray , roman_s , roman_m end_POSTSUBSCRIPT caligraphic_T start_POSTSUBSCRIPT italic_λ , roman_m end_POSTSUBSCRIPT ( italic_t ) italic_I start_POSTSUBSCRIPT italic_λ , roman_s , roman_m end_POSTSUBSCRIPT (A15)
=δray,s𝒯λ(t)¯Iλ,s,absent¯subscript𝛿rayssubscript𝒯𝜆𝑡subscript𝐼𝜆s\displaystyle=\overline{\delta_{\mathrm{ray,s}}\mathcal{T}_{\mathrm{\lambda}}(% t)}I_{\mathrm{\lambda,s}},= over¯ start_ARG italic_δ start_POSTSUBSCRIPT roman_ray , roman_s end_POSTSUBSCRIPT caligraphic_T start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_t ) end_ARG italic_I start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT ,

where

δray,s={1,\textifrayintersectsstar.0,\textelse.subscript𝛿rayscases1\text𝑖𝑓𝑟𝑎𝑦𝑖𝑛𝑡𝑒𝑟𝑠𝑒𝑐𝑡𝑠𝑠𝑡𝑎𝑟0\text𝑒𝑙𝑠𝑒\delta_{\mathrm{ray,s}}=\left\{\begin{array}[]{ll}1,&\text{ifrayintersectsstar% .}\\ 0,&\text{else.}\end{array}\right.italic_δ start_POSTSUBSCRIPT roman_ray , roman_s end_POSTSUBSCRIPT = { start_ARRAY start_ROW start_CELL 1 , end_CELL start_CELL italic_i italic_f italic_r italic_a italic_y italic_i italic_n italic_t italic_e italic_r italic_s italic_e italic_c italic_t italic_s italic_s italic_t italic_a italic_r . end_CELL end_ROW start_ROW start_CELL 0 , end_CELL start_CELL italic_e italic_l italic_s italic_e . end_CELL end_ROW end_ARRAY (A16)

and

𝒯λ,m(t)=eτλ,path,m(t),subscript𝒯𝜆m𝑡superscript𝑒subscript𝜏𝜆pathm𝑡\displaystyle\mathcal{T}_{\mathrm{\lambda,m}}(t)=e^{-\tau_{\mathrm{\lambda,% path,m}}(t)},caligraphic_T start_POSTSUBSCRIPT italic_λ , roman_m end_POSTSUBSCRIPT ( italic_t ) = italic_e start_POSTSUPERSCRIPT - italic_τ start_POSTSUBSCRIPT italic_λ , roman_path , roman_m end_POSTSUBSCRIPT ( italic_t ) end_POSTSUPERSCRIPT ,
τλ,path,m(t)=0α~λ(sm,t)𝑑sm.subscript𝜏𝜆pathm𝑡superscriptsubscript0subscript~𝛼𝜆subscript𝑠𝑚𝑡differential-dsubscript𝑠𝑚\displaystyle\tau_{\mathrm{\lambda,path,m}}(t)=\int_{0}^{\infty}\tilde{\alpha}% _{\lambda}\left(s_{m},t\right)ds_{m}.italic_τ start_POSTSUBSCRIPT italic_λ , roman_path , roman_m end_POSTSUBSCRIPT ( italic_t ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT over~ start_ARG italic_α end_ARG start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_s start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT , italic_t ) italic_d italic_s start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT . (A17)

Here, we neglect stellar limb-darkening, making Iλ,sI¯λ,ssubscript𝐼𝜆𝑠subscript¯𝐼𝜆𝑠\frac{I_{\lambda,s}}{\overline{I}_{\lambda,s}}divide start_ARG italic_I start_POSTSUBSCRIPT italic_λ , italic_s end_POSTSUBSCRIPT end_ARG start_ARG over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , italic_s end_POSTSUBSCRIPT end_ARG to be factored out of the integrals in Equation (A14). Therefore,

Δλ(t)=subscriptΔ𝜆𝑡absent\displaystyle\Delta_{\lambda}(t)=roman_Δ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_t ) = [Iλ,sI¯λ,sAoverlap(t)Iλ,sI¯λ,sAp(full)δray,s𝒯λ(t)¯𝑑A+I¯λ,p(night)I¯λ,sAp(full)Ap(full)V(θ,ϕ,t)Iλ,p(night)(θ,ϕ,t)I¯λ,s𝑑A]delimited-[]subscript𝐼𝜆ssubscript¯𝐼𝜆ssubscript𝐴overlap𝑡subscript𝐼𝜆ssubscript¯𝐼𝜆ssubscriptsubscript𝐴pfull¯subscript𝛿rayssubscript𝒯𝜆𝑡differential-d𝐴subscript¯𝐼𝜆pnightsubscript¯𝐼𝜆ssubscript𝐴pfullsubscriptsubscript𝐴pfull𝑉𝜃italic-ϕ𝑡subscript𝐼𝜆pnight𝜃italic-ϕ𝑡subscript¯𝐼𝜆𝑠differential-d𝐴\displaystyle\left[\frac{I_{\mathrm{\lambda,s}}}{\overline{I}_{\mathrm{\lambda% ,s}}}A_{\mathrm{overlap}}(t)-\frac{I_{\mathrm{\lambda,s}}}{\overline{I}_{% \mathrm{\lambda,s}}}\int_{A_{\mathrm{p(full)}}}\overline{\delta_{\mathrm{ray,s% }}\mathcal{T}_{\lambda}(t)}dA+\frac{\overline{I}_{\mathrm{\lambda,p(night)}}}{% \overline{I}_{\mathrm{\lambda,s}}}A_{\mathrm{p(full)}}-\int_{A_{\mathrm{p(full% )}}}\frac{V(\theta,\phi,t)I_{\mathrm{\lambda,p(night)}}(\theta,\phi,t)}{% \overline{I}_{\lambda,s}}dA\right][ divide start_ARG italic_I start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT end_ARG start_ARG over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT end_ARG italic_A start_POSTSUBSCRIPT roman_overlap end_POSTSUBSCRIPT ( italic_t ) - divide start_ARG italic_I start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT end_ARG start_ARG over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT end_ARG ∫ start_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT roman_p ( roman_full ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT over¯ start_ARG italic_δ start_POSTSUBSCRIPT roman_ray , roman_s end_POSTSUBSCRIPT caligraphic_T start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_t ) end_ARG italic_d italic_A + divide start_ARG over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT end_ARG start_ARG over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT end_ARG italic_A start_POSTSUBSCRIPT roman_p ( roman_full ) end_POSTSUBSCRIPT - ∫ start_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT roman_p ( roman_full ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT divide start_ARG italic_V ( italic_θ , italic_ϕ , italic_t ) italic_I start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT ( italic_θ , italic_ϕ , italic_t ) end_ARG start_ARG over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , italic_s end_POSTSUBSCRIPT end_ARG italic_d italic_A ] (A18)
×1As(full)+I¯λ,p(night)I¯λ,sAp(full),absent1subscript𝐴sfullsubscript¯𝐼𝜆pnightsubscript¯𝐼𝜆ssubscript𝐴pfull\displaystyle\times\frac{1}{A_{\mathrm{s(full)}}+\frac{\overline{I}_{\mathrm{% \lambda,p(night)}}}{\overline{I}_{\mathrm{\lambda,s}}}A_{\mathrm{p(full)}}},× divide start_ARG 1 end_ARG start_ARG italic_A start_POSTSUBSCRIPT roman_s ( roman_full ) end_POSTSUBSCRIPT + divide start_ARG over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT end_ARG start_ARG over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT end_ARG italic_A start_POSTSUBSCRIPT roman_p ( roman_full ) end_POSTSUBSCRIPT end_ARG ,

where Aoverlapsubscript𝐴overlapA_{\mathrm{overlap}}italic_A start_POSTSUBSCRIPT roman_overlap end_POSTSUBSCRIPT denotes the overlapping area between the planetary and stellar disks (see MacDonald & Lewis, 2022). Then, we introduce two factors:

ελ,hetsubscript𝜀𝜆het\displaystyle\varepsilon_{\mathrm{\lambda,het}}italic_ε start_POSTSUBSCRIPT italic_λ , roman_het end_POSTSUBSCRIPT Iλ,sI¯λ,s=11i=1Nhetfhet,i(1Iλ,het,iIλ,s),absentsubscript𝐼𝜆ssubscript¯𝐼𝜆s11superscriptsubscript𝑖1subscript𝑁hetsubscript𝑓heti1subscript𝐼𝜆hetisubscript𝐼𝜆s\displaystyle\equiv\frac{I_{\mathrm{\lambda,s}}}{\overline{I}_{\mathrm{\lambda% ,s}}}=\frac{1}{1-\sum_{i=1}^{N_{\mathrm{het}}}f_{\mathrm{het,i}}\left(1-\frac{% I_{\mathrm{\lambda,het,i}}}{I_{\mathrm{\lambda,s}}}\right)},≡ divide start_ARG italic_I start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT end_ARG start_ARG over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT end_ARG = divide start_ARG 1 end_ARG start_ARG 1 - ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT roman_het end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_f start_POSTSUBSCRIPT roman_het , roman_i end_POSTSUBSCRIPT ( 1 - divide start_ARG italic_I start_POSTSUBSCRIPT italic_λ , roman_het , roman_i end_POSTSUBSCRIPT end_ARG start_ARG italic_I start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT end_ARG ) end_ARG , (A19)
ελ,nightsubscript𝜀𝜆night\displaystyle\varepsilon_{\mathrm{\lambda,night}}italic_ε start_POSTSUBSCRIPT italic_λ , roman_night end_POSTSUBSCRIPT Iλ,p(night)I¯λ,s.absentsubscript𝐼𝜆pnightsubscript¯𝐼𝜆s\displaystyle\equiv\frac{I_{\mathrm{\lambda,p(night)}}}{\overline{I}_{\mathrm{% \lambda,s}}}.≡ divide start_ARG italic_I start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT end_ARG start_ARG over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT end_ARG .

Therefore, substituting Equation (A19) into Equation (A18), we have

Δλ(t)={ελ,het[Aoverlap(t)Ap(t)δray,s𝒯λ(t)¯𝑑A]+ε¯λ,nightAp(full)Ap(full)V(θ,ϕ,t)Iλ,p(night)(θ,ϕ,t)I¯λ,s𝑑A}×1As(full)+ε¯λ,nightAp(full).subscriptΔ𝜆𝑡subscript𝜀𝜆hetdelimited-[]subscript𝐴overlap𝑡subscriptsubscript𝐴p𝑡¯subscript𝛿rayssubscript𝒯𝜆𝑡differential-d𝐴subscript¯𝜀𝜆nightsubscript𝐴pfullsubscriptsubscript𝐴pfull𝑉𝜃italic-ϕ𝑡subscript𝐼𝜆pnight𝜃italic-ϕ𝑡subscript¯𝐼𝜆𝑠differential-d𝐴1subscript𝐴sfullsubscript¯𝜀𝜆nightsubscript𝐴pfull\Delta_{\lambda}(t)=\left\{\varepsilon_{\mathrm{\lambda,het}}\left[A_{\mathrm{% overlap}}(t)-\int_{A_{\mathrm{p}}(t)}\overline{\delta_{\mathrm{ray,s}}\mathcal% {T}_{\lambda}(t)}dA\right]+\overline{\varepsilon}_{\mathrm{\lambda,night}}A_{% \mathrm{p(full)}}-\int_{A_{\mathrm{p(full)}}}\frac{V(\theta,\phi,t)I_{\mathrm{% \lambda,p(night)}}(\theta,\phi,t)}{\bar{I}_{\lambda,s}}dA\right\}\times\frac{1% }{A_{\mathrm{s(full)}}+\overline{\varepsilon}_{\mathrm{\lambda,night}}A_{% \mathrm{p(full)}}}.roman_Δ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_t ) = { italic_ε start_POSTSUBSCRIPT italic_λ , roman_het end_POSTSUBSCRIPT [ italic_A start_POSTSUBSCRIPT roman_overlap end_POSTSUBSCRIPT ( italic_t ) - ∫ start_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT ( italic_t ) end_POSTSUBSCRIPT over¯ start_ARG italic_δ start_POSTSUBSCRIPT roman_ray , roman_s end_POSTSUBSCRIPT caligraphic_T start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_t ) end_ARG italic_d italic_A ] + over¯ start_ARG italic_ε end_ARG start_POSTSUBSCRIPT italic_λ , roman_night end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT roman_p ( roman_full ) end_POSTSUBSCRIPT - ∫ start_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT roman_p ( roman_full ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT divide start_ARG italic_V ( italic_θ , italic_ϕ , italic_t ) italic_I start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT ( italic_θ , italic_ϕ , italic_t ) end_ARG start_ARG over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , italic_s end_POSTSUBSCRIPT end_ARG italic_d italic_A } × divide start_ARG 1 end_ARG start_ARG italic_A start_POSTSUBSCRIPT roman_s ( roman_full ) end_POSTSUBSCRIPT + over¯ start_ARG italic_ε end_ARG start_POSTSUBSCRIPT italic_λ , roman_night end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT roman_p ( roman_full ) end_POSTSUBSCRIPT end_ARG . (A20)

The emergent planetary thermal emission intensity Iλ,p(night)subscript𝐼𝜆pnightI_{\mathrm{\lambda,p(night)}}italic_I start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT can be solved by the radiative transfer equation (e.g., two-stream radiative transfer method). Its solution depends on the source function and optical depth, while the visible function depends on time, from which it follows that the flux depends strongly on time (orbital phase), reflecting the importance of the 3D atmospheric structure and radiative transfer calculations for transmission spectroscopy.

Now let’s start making approximations. The 3D effects are not the focus of this paper and we simplify to 1D atmospheres:

Iλ,p(night)(θ,ϕ,t)=Iλ,p(night)(t),subscript𝐼𝜆pnight𝜃italic-ϕ𝑡subscript𝐼𝜆pnight𝑡\displaystyle I_{\mathrm{\lambda,p(night)}}(\theta,\phi,t)=I_{\mathrm{\lambda,% p(night)}}(t),italic_I start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT ( italic_θ , italic_ϕ , italic_t ) = italic_I start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT ( italic_t ) ,
V(θ,ϕ,t)=1.𝑉𝜃italic-ϕ𝑡1\displaystyle V(\theta,\phi,t)=1.italic_V ( italic_θ , italic_ϕ , italic_t ) = 1 . (A21)

Therefore,

Ap(full)V(θ,ϕ,t)Iλ,p(night)(θ,ϕ,t)I¯λ,s𝑑A=Iλ,p(night)(t)I¯λ,sAp(full)𝑑A=ελ,night(t)Ap(full).subscriptsubscript𝐴pfull𝑉𝜃italic-ϕ𝑡subscript𝐼𝜆pnight𝜃italic-ϕ𝑡subscript¯𝐼𝜆𝑠differential-d𝐴subscript𝐼𝜆pnight𝑡subscript¯𝐼𝜆ssubscriptsubscript𝐴pfulldifferential-d𝐴subscript𝜀𝜆night𝑡subscript𝐴pfull\int_{A_{\mathrm{p(full)}}}\frac{V(\theta,\phi,t)I_{\mathrm{\lambda,p(night)}}% (\theta,\phi,t)}{\overline{I}_{\lambda,s}}dA=\frac{I_{\mathrm{\lambda,p(night)% }}(t)}{\overline{I}_{\mathrm{\lambda,s}}}\int_{A_{\mathrm{p(full)}}}dA=% \varepsilon_{\mathrm{\lambda,night}}(t)A_{\mathrm{p(full)}}.∫ start_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT roman_p ( roman_full ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT divide start_ARG italic_V ( italic_θ , italic_ϕ , italic_t ) italic_I start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT ( italic_θ , italic_ϕ , italic_t ) end_ARG start_ARG over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , italic_s end_POSTSUBSCRIPT end_ARG italic_d italic_A = divide start_ARG italic_I start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT ( italic_t ) end_ARG start_ARG over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT end_ARG ∫ start_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT roman_p ( roman_full ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_d italic_A = italic_ε start_POSTSUBSCRIPT italic_λ , roman_night end_POSTSUBSCRIPT ( italic_t ) italic_A start_POSTSUBSCRIPT roman_p ( roman_full ) end_POSTSUBSCRIPT . (A22)

Then, we introduce

Aλ,eff(t)Aoverlap(t)Ap(full)δray,s𝒯λ(t)¯𝑑A=πRλ,eff2(t),subscript𝐴𝜆eff𝑡subscript𝐴overlap𝑡subscriptsubscript𝐴pfull¯subscript𝛿rayssubscript𝒯𝜆𝑡differential-d𝐴𝜋superscriptsubscript𝑅𝜆eff2𝑡\displaystyle A_{\mathrm{\lambda,eff}}(t)\equiv A_{\mathrm{overlap}}(t)-\int_{% A_{\mathrm{p(full)}}}\overline{\delta_{\mathrm{ray,s}}\mathcal{T}_{\lambda}(t)% }dA=\pi R_{\mathrm{\lambda,eff}}^{2}(t),italic_A start_POSTSUBSCRIPT italic_λ , roman_eff end_POSTSUBSCRIPT ( italic_t ) ≡ italic_A start_POSTSUBSCRIPT roman_overlap end_POSTSUBSCRIPT ( italic_t ) - ∫ start_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT roman_p ( roman_full ) end_POSTSUBSCRIPT end_POSTSUBSCRIPT over¯ start_ARG italic_δ start_POSTSUBSCRIPT roman_ray , roman_s end_POSTSUBSCRIPT caligraphic_T start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_t ) end_ARG italic_d italic_A = italic_π italic_R start_POSTSUBSCRIPT italic_λ , roman_eff end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_t ) ,
dλ,night11+¯λ,p(night),out(t)λ,s,out=11+Ap(full)I¯λ,p(night)As(full)I¯λ,s=As(full)As(full)+Ap(full)ε¯λ,night.subscript𝑑𝜆night11subscript¯𝜆pnightout𝑡subscript𝜆sout11subscript𝐴pfullsubscript¯𝐼𝜆pnightsubscript𝐴sfullsubscript¯𝐼𝜆ssubscript𝐴sfullsubscript𝐴sfullsubscript𝐴pfullsubscript¯𝜀𝜆night\displaystyle d_{\mathrm{\lambda,night}}\equiv\frac{1}{1+\frac{\overline{% \mathcal{F}}_{\mathrm{\lambda,p(night),out}}(t)}{\mathcal{F}_{\mathrm{\lambda,% s,out}}}}=\frac{1}{1+\frac{A_{\mathrm{p(full)}}\overline{I}_{\mathrm{\lambda,p% (night)}}}{A_{\mathrm{s(full)}}\overline{I}_{\mathrm{\lambda,s}}}}=\frac{A_{% \mathrm{s(full)}}}{A_{\mathrm{s(full)}}+A_{\mathrm{p(full)}}\overline{% \varepsilon}_{\mathrm{\lambda,night}}}.italic_d start_POSTSUBSCRIPT italic_λ , roman_night end_POSTSUBSCRIPT ≡ divide start_ARG 1 end_ARG start_ARG 1 + divide start_ARG over¯ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) , roman_out end_POSTSUBSCRIPT ( italic_t ) end_ARG start_ARG caligraphic_F start_POSTSUBSCRIPT italic_λ , roman_s , roman_out end_POSTSUBSCRIPT end_ARG end_ARG = divide start_ARG 1 end_ARG start_ARG 1 + divide start_ARG italic_A start_POSTSUBSCRIPT roman_p ( roman_full ) end_POSTSUBSCRIPT over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT end_ARG start_ARG italic_A start_POSTSUBSCRIPT roman_s ( roman_full ) end_POSTSUBSCRIPT over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT end_ARG end_ARG = divide start_ARG italic_A start_POSTSUBSCRIPT roman_s ( roman_full ) end_POSTSUBSCRIPT end_ARG start_ARG italic_A start_POSTSUBSCRIPT roman_s ( roman_full ) end_POSTSUBSCRIPT + italic_A start_POSTSUBSCRIPT roman_p ( roman_full ) end_POSTSUBSCRIPT over¯ start_ARG italic_ε end_ARG start_POSTSUBSCRIPT italic_λ , roman_night end_POSTSUBSCRIPT end_ARG . (A23)

Substituting Equation (A22) and (A) into Equation (A20) and omitting “full”, we have

Δλ(t)={ελ,hetAλ,eff(t)As+ApAs[ε¯λ,nightελ,night(t)]}×dλ,night.subscriptΔ𝜆𝑡subscript𝜀𝜆hetsubscript𝐴𝜆eff𝑡subscript𝐴ssubscript𝐴psubscript𝐴sdelimited-[]subscript¯𝜀𝜆nightsubscript𝜀𝜆night𝑡subscript𝑑𝜆night\Delta_{\lambda}(t)=\left\{\varepsilon_{\mathrm{\lambda,het}}\frac{A_{\mathrm{% \lambda,eff}}(t)}{A_{\mathrm{s}}}+\frac{A_{\mathrm{p}}}{A_{\mathrm{s}}}\left[% \bar{\varepsilon}_{\mathrm{\lambda,night}}-\varepsilon_{\mathrm{\lambda,night}% }(t)\right]\right\}\times d_{\mathrm{\lambda,night}}.roman_Δ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_t ) = { italic_ε start_POSTSUBSCRIPT italic_λ , roman_het end_POSTSUBSCRIPT divide start_ARG italic_A start_POSTSUBSCRIPT italic_λ , roman_eff end_POSTSUBSCRIPT ( italic_t ) end_ARG start_ARG italic_A start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT end_ARG + divide start_ARG italic_A start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT end_ARG start_ARG italic_A start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT end_ARG [ over¯ start_ARG italic_ε end_ARG start_POSTSUBSCRIPT italic_λ , roman_night end_POSTSUBSCRIPT - italic_ε start_POSTSUBSCRIPT italic_λ , roman_night end_POSTSUBSCRIPT ( italic_t ) ] } × italic_d start_POSTSUBSCRIPT italic_λ , roman_night end_POSTSUBSCRIPT . (A24)

Taking into account the photospheric radius correction (Fortney et al., 2019), Ap=Aλ,eff(t)subscript𝐴psubscript𝐴𝜆eff𝑡A_{\mathrm{p}}=A_{\mathrm{\lambda,eff}}(t)italic_A start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT = italic_A start_POSTSUBSCRIPT italic_λ , roman_eff end_POSTSUBSCRIPT ( italic_t ), we have,

Δλ(t)={ελ,hetAλ,eff(t)As+Aλ,eff(t)As[ε¯λ,nightελ,night(t)]}×dλ,night.subscriptΔ𝜆𝑡subscript𝜀𝜆hetsubscript𝐴𝜆eff𝑡subscript𝐴ssubscript𝐴𝜆eff𝑡subscript𝐴sdelimited-[]subscript¯𝜀𝜆nightsubscript𝜀𝜆night𝑡subscript𝑑𝜆night\Delta_{\lambda}(t)=\left\{\varepsilon_{\mathrm{\lambda,het}}\frac{A_{\mathrm{% \lambda,eff}}(t)}{A_{\mathrm{s}}}+\frac{A_{\mathrm{\lambda,eff}}(t)}{A_{% \mathrm{s}}}\left[\bar{\varepsilon}_{\mathrm{\lambda,night}}-\varepsilon_{% \mathrm{\lambda,night}}(t)\right]\right\}\times d_{\mathrm{\lambda,night}}.roman_Δ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_t ) = { italic_ε start_POSTSUBSCRIPT italic_λ , roman_het end_POSTSUBSCRIPT divide start_ARG italic_A start_POSTSUBSCRIPT italic_λ , roman_eff end_POSTSUBSCRIPT ( italic_t ) end_ARG start_ARG italic_A start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT end_ARG + divide start_ARG italic_A start_POSTSUBSCRIPT italic_λ , roman_eff end_POSTSUBSCRIPT ( italic_t ) end_ARG start_ARG italic_A start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT end_ARG [ over¯ start_ARG italic_ε end_ARG start_POSTSUBSCRIPT italic_λ , roman_night end_POSTSUBSCRIPT - italic_ε start_POSTSUBSCRIPT italic_λ , roman_night end_POSTSUBSCRIPT ( italic_t ) ] } × italic_d start_POSTSUBSCRIPT italic_λ , roman_night end_POSTSUBSCRIPT . (A25)

Alternatively, we introduce an approximate line intensity L𝐿Litalic_L,

Lλ,p(night)(t)subscript𝐿𝜆pnight𝑡\displaystyle L_{\mathrm{\lambda,p(night)}}(t)italic_L start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT ( italic_t ) Iλ,p(night)(t)I¯λ,p(night)absentsubscript𝐼𝜆pnight𝑡subscript¯𝐼𝜆pnight\displaystyle\equiv I_{\mathrm{\lambda,p(night)}}(t)-\overline{I}_{\mathrm{% \lambda,p(night)}}≡ italic_I start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT ( italic_t ) - over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT (A26)
Iλ,p(night)(t)I¯λ,p(night),continuum.absentsubscript𝐼𝜆pnight𝑡subscript¯𝐼𝜆pnightcontinuum\displaystyle\approx I_{\mathrm{\lambda,p(night)}}(t)-\overline{I}_{\mathrm{% \lambda,p(night),continuum}}.≈ italic_I start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT ( italic_t ) - over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) , roman_continuum end_POSTSUBSCRIPT .

Substituting Equation (A26) into Equation (A25), we have

Δλ(t)=[ελ,hetAλ,eff(t)AsLλ,p(night)(t)I¯λ,sAλ,eff(t)As]×dλ,night,subscriptΔ𝜆𝑡delimited-[]subscript𝜀𝜆hetsubscript𝐴𝜆eff𝑡subscript𝐴ssubscript𝐿𝜆pnight𝑡subscript¯𝐼𝜆ssubscript𝐴𝜆eff𝑡subscript𝐴ssubscript𝑑𝜆night\Delta_{\lambda}(t)=\left[\varepsilon_{\mathrm{\lambda,het}}\frac{A_{\mathrm{% \lambda,eff}}(t)}{A_{\mathrm{s}}}-\frac{L_{\mathrm{\lambda,p(night)}}(t)}{% \overline{I}_{\mathrm{\lambda,s}}}\frac{A_{\mathrm{\lambda,eff}}(t)}{A_{% \mathrm{s}}}\right]\times d_{\mathrm{\lambda,night}},roman_Δ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_t ) = [ italic_ε start_POSTSUBSCRIPT italic_λ , roman_het end_POSTSUBSCRIPT divide start_ARG italic_A start_POSTSUBSCRIPT italic_λ , roman_eff end_POSTSUBSCRIPT ( italic_t ) end_ARG start_ARG italic_A start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT end_ARG - divide start_ARG italic_L start_POSTSUBSCRIPT italic_λ , roman_p ( roman_night ) end_POSTSUBSCRIPT ( italic_t ) end_ARG start_ARG over¯ start_ARG italic_I end_ARG start_POSTSUBSCRIPT italic_λ , roman_s end_POSTSUBSCRIPT end_ARG divide start_ARG italic_A start_POSTSUBSCRIPT italic_λ , roman_eff end_POSTSUBSCRIPT ( italic_t ) end_ARG start_ARG italic_A start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT end_ARG ] × italic_d start_POSTSUBSCRIPT italic_λ , roman_night end_POSTSUBSCRIPT , (A27)

which completes the derivation of Equation (4), our unified equation for high-resolution transmission spectra with 1D atmosphere.

Next, we further analyze this equation in order to discuss what factors influence the intensity of the transmission spectrum. The formal solution for 1D plane-parallel radiative transfer equation (without scattering and μ=1𝜇1\mu=1italic_μ = 1) is

Iλ=Iλboteτλ,vert\textbot+0τλ,vert\textbotSλ(τλ,vert)eτλ,vert𝑑τλ,vert,subscript𝐼𝜆superscriptsubscript𝐼𝜆botsuperscript𝑒superscriptsubscript𝜏𝜆vert\text𝑏𝑜𝑡superscriptsubscript0superscriptsubscript𝜏𝜆vert\text𝑏𝑜𝑡subscript𝑆𝜆subscript𝜏𝜆vertsuperscript𝑒subscript𝜏𝜆vertdifferential-dsubscript𝜏𝜆vertI_{\mathrm{\lambda}}=I_{\lambda}^{\mathrm{bot}}e^{-\tau_{\mathrm{\lambda,vert}% }^{\text{bot}}}+\int_{0}^{\tau_{\mathrm{\lambda,vert}}^{\text{bot}}}S_{\lambda% }(\tau_{\mathrm{\lambda,vert}})e^{-\tau_{\mathrm{\lambda,vert}}}d\tau_{\mathrm% {\lambda,vert}},italic_I start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT = italic_I start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_bot end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_τ start_POSTSUBSCRIPT italic_λ , roman_vert end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_b italic_o italic_t end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT + ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_τ start_POSTSUBSCRIPT italic_λ , roman_vert end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_b italic_o italic_t end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT italic_S start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_τ start_POSTSUBSCRIPT italic_λ , roman_vert end_POSTSUBSCRIPT ) italic_e start_POSTSUPERSCRIPT - italic_τ start_POSTSUBSCRIPT italic_λ , roman_vert end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_d italic_τ start_POSTSUBSCRIPT italic_λ , roman_vert end_POSTSUBSCRIPT , (A28)

where Sλ(τ)subscript𝑆𝜆𝜏S_{\mathrm{\lambda}}(\tau)italic_S start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_τ ) is source function, τ𝜏\tauitalic_τ is optical depth. Similarly, we introduce

𝒯λ,verteτλ,vert,subscript𝒯𝜆vertsuperscript𝑒subscript𝜏𝜆vert\mathcal{T}_{\mathrm{\lambda,vert}}\equiv e^{-\tau_{\mathrm{\lambda,vert}}},caligraphic_T start_POSTSUBSCRIPT italic_λ , roman_vert end_POSTSUBSCRIPT ≡ italic_e start_POSTSUPERSCRIPT - italic_τ start_POSTSUBSCRIPT italic_λ , roman_vert end_POSTSUBSCRIPT end_POSTSUPERSCRIPT , (A29)

which denotes the atmospheric transmission from a given layer to the top of the atmosphere. Substituting Equation (A29) into Equation (A28), we have

Iλ(μ)=Iλbot𝒯λ,vert,bot+𝒯λ,vert,bot1Sλ(𝒯λ,vert)𝑑𝒯λ,vert,subscript𝐼𝜆𝜇superscriptsubscript𝐼𝜆botsubscript𝒯𝜆vertbotsuperscriptsubscriptsubscript𝒯𝜆vertbot1subscript𝑆𝜆subscript𝒯𝜆vertdifferential-dsubscript𝒯𝜆vertI_{\mathrm{\lambda}}(\mu)=I_{\mathrm{\lambda}}^{\mathrm{bot}}\mathcal{T}_{% \mathrm{\lambda,vert,bot}}+\int_{\mathcal{T}_{\lambda,\mathrm{vert,bot}}}^{1}S% _{\mathrm{\lambda}}(\mathcal{T}_{\mathrm{\lambda,vert}})d\mathcal{T}_{\mathrm{% \lambda,vert}},italic_I start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( italic_μ ) = italic_I start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_bot end_POSTSUPERSCRIPT caligraphic_T start_POSTSUBSCRIPT italic_λ , roman_vert , roman_bot end_POSTSUBSCRIPT + ∫ start_POSTSUBSCRIPT caligraphic_T start_POSTSUBSCRIPT italic_λ , roman_vert , roman_bot end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT italic_S start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( caligraphic_T start_POSTSUBSCRIPT italic_λ , roman_vert end_POSTSUBSCRIPT ) italic_d caligraphic_T start_POSTSUBSCRIPT italic_λ , roman_vert end_POSTSUBSCRIPT , (A30)

where the core is to solve for Sλsubscript𝑆𝜆S_{\mathrm{\lambda}}italic_S start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT.

To show L𝐿Litalic_L, next, we introduce Milne-Eddington approximation,

τλ=kl+kckc𝑑τc,subscript𝜏𝜆subscript𝑘𝑙subscript𝑘𝑐subscript𝑘𝑐differential-dsubscript𝜏𝑐\tau_{\lambda}=\int\frac{k_{l}+k_{c}}{k_{c}}d\tau_{c},italic_τ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT = ∫ divide start_ARG italic_k start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT + italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG start_ARG italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG italic_d italic_τ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT , (A31)

where

klkc=ηλ(constant),S(τλ)=S0+dSdτcτc,formulae-sequencesubscript𝑘𝑙subscript𝑘𝑐subscript𝜂𝜆constant𝑆subscript𝜏𝜆subscript𝑆0𝑑𝑆𝑑subscript𝜏𝑐subscript𝜏𝑐\frac{k_{l}}{k_{c}}=\eta_{\mathrm{\lambda}}\mathrm{(constant)},\quad S\left(% \tau_{\lambda}\right)=S_{\mathrm{0}}+\frac{dS}{d\tau_{c}}\tau_{c},divide start_ARG italic_k start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT end_ARG start_ARG italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG = italic_η start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ( roman_constant ) , italic_S ( italic_τ start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ) = italic_S start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + divide start_ARG italic_d italic_S end_ARG start_ARG italic_d italic_τ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG italic_τ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT , (A32)

where klsubscript𝑘𝑙k_{l}italic_k start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT and kcsubscript𝑘ck_{\mathrm{c}}italic_k start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT are line absorption coefficient and continuum absorption coefficient, respectively. In local thermodynamic equilibrium, S=B(Tvert)𝑆𝐵subscript𝑇vertS=B(T_{\mathrm{vert}})italic_S = italic_B ( italic_T start_POSTSUBSCRIPT roman_vert end_POSTSUBSCRIPT ).

Substituting Equation (A30)-(A32) into Equation (A26), therefore,

L=dB(Tvert)dτc(11+ηλ1).𝐿𝑑𝐵subscript𝑇vert𝑑subscript𝜏𝑐11subscript𝜂𝜆1L=\frac{dB(T_{\mathrm{vert}})}{d\tau_{c}}\cdot\left(\frac{1}{1+\eta_{\lambda}}% -1\right).italic_L = divide start_ARG italic_d italic_B ( italic_T start_POSTSUBSCRIPT roman_vert end_POSTSUBSCRIPT ) end_ARG start_ARG italic_d italic_τ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG ⋅ ( divide start_ARG 1 end_ARG start_ARG 1 + italic_η start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT end_ARG - 1 ) . (A33)

Appendix B A Detailed Description of Data Reduction

The observation logs for the archival data we used are shown in Table 1. The raw CARMENES data were reduced using the CARACAL pipeline (Caballero et al., 2016). The GIANO-B observations were performed in ABAB nodding mode, which allows efficient removal of sky background and detector noise. The raw spectra were reduced using the GOFIO pipeline (Rainer et al., 2018). The pipeline produces three spectra, including nodding A, nodding B, and combined nodding AB. In this work, only the AB spectra are used. Although the GOFIO pipeline performed a preliminary wavelength calibration using the U-Ne lamp spectra as a template, the mechanical instability of GIANO-B made the wavelength solution unstable in the observations. To prevent saturation of some emission lines from contaminating the observations, the U-Ne lamp spectra were acquired only at the end of the observations, resulting in an inaccurate wavelength solution determined by GOFIO (Brogi et al., 2018). Therefore, we realigned the spectra and performed wavelength corrections using telluric lines following the approach of Guilluy et al. (2019), Giacobbe et al. (2021), and Yan et al. (2022b). After raw data reduction by the pipelines, both CARMENES and GIANO-B data are archived as order-by-order spectra at the vacuum wavelength and in the terrestrial rest frame.

Table 1: Summary of the Observation Logs.
Instrument Target Night Date [UT] Exposure Time [s] Nobssubscript𝑁obsN_{\mathrm{obs}}italic_N start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT Airmass
GIANO-B WASP-33b Night-G1 2018-10-17 200 167 1.66-1.01-1.49
GIANO-B WASP-33b Night-G2 2018-11-08 200 190 1.77-1.01-1.68
CARMENES WASP-33b Night-C1 2017-01-05 120 94 1.00-1.00-1.52
CARMENES WASP-33b Night-C2 2018-11-30 250 31 1.15-1.00-1.08
CARMENES WASP-33b Night-C3 2018-12-05 200 64 1.39-1.00-1.14
CARMENES WASP-33b Night-C4 2019-10-01 120 101 2.04-1.15-1.00
CARMENES WASP-33b Night-C5 2019-11-03 210 45 1.75-1.26-1.06

We adopt the method detailed by Yan et al. (2022a) and Yang et al. (2024) to perform the subsequent spectral data reduction. To correct for outliers that are unlikely to be caused by the observed source, we flagged the pixels that were greater than five standard deviations from the flux values of the smoothed spectrum produced by a median filter with a window size of nine pixels for each order. The flux values of these pixels were replaced by the median of the thirty surrounding pixels (Langeveld et al., 2021). To correct for continuum variation, we divided each spectrum by the best-fit polynomial function, which is derived by fitting the ratio of each individual spectrum to the reference spectrum (i.e., the average of all spectra) with a fourth-order polynomial function (Chen et al., 2020; Yang et al., 2024). We constructed the spectral matrix by collecting all spectra of each observation, sorted by time. For GIANO-B, we discarded orders 8, 9, 10, 23, 24, 30, 45, 46, 47, 48, and 49 from our analysis, where order 0 is the reddest and order 49 is the bluest, due to saturated telluric absorption at certain wavelengths or few telluric lines to obtain wavelength corrections, similar to Giacobbe et al. (2021). For CARMENES, we discard orders 6, 7, 8, 9, and 10, where order 0 is the reddest and order 27 is the bluest. The discarded orders are shown in Figure 6.

Refer to caption
Figure 6: Masking of the CARMENES and GIANO-B spectral orders. The vertical red and gray shaded areas indicate the masked orders for CARMENES and GIANO-B, respectively. The green line indicates the H2O thermal emission spectrum. The purple line indicates the Earth’s telluric spectrum.

To remove strong telluric regions, we masked the deep telluric regions that fall below 30% of the median of the master spectrum, which is calculated by averaging all spectra for each order and contains the region around the deep telluric lines below 95% of the median of the master spectrum. To remove strong sky emission lines, we masked the region above 150% of the median of the continuum normalized master spectrum (containing the region above 105% around each sky emission line). In addition, we further removed the 4σ𝜎\sigmaitalic_σ outliers in each wavelength channel (i.e., each column) using an iterative mode. Finally, we normalized each spectrum with a Gaussian high-pass filter. We used the SYSREM algorithm (Tamuz et al., 2005) to reduce the telluric and stellar lines in the terrestrial rest frame. The SYSREM algorithm is used to fit each wavelength channel (each column) and each spectrum (each row) in the spectral matrix to capture systematics due to telluric contamination, stellar lines, and instrumental effects. These systematics are then removed from the spectral matrix. The SYSREM iteratively refines this process, yielding residual spectral matrices that retain planetary features and noise. We followed the approach proposed by Gibson et al. (2020), which divides the spectral matrix by the systematics rather than subtracting them. We ran the algorithm for fifteen iterations and selected the residual spectral matrix that yielded the maximum detection significance as the final output.

Appendix C Individual KpΔ\varvsubscript𝐾pΔ\varvK_{\mathrm{p}}-\Delta\varvitalic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT - roman_Δ map for each night

The individual KpΔ\varvsubscript𝐾pΔ\varvK_{\mathrm{p}}-\Delta\varvitalic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT - roman_Δ maps for each GIANO-B observation are shown in Figure 7.

Refer to caption
Refer to caption
Figure 7: KpΔ\varvsubscript𝐾pΔ\varvK_{\mathrm{p}}-\Delta\varvitalic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT - roman_Δ maps of H2O from individual nights of GIANO-B observations. The crossed white dashed lines indicate the S/N peak.

Appendix D Truncated GIANO-B data to resemble the insufficient wavelength coverage of CARMENES

We truncated the GIANO-B data at less-than-or-similar-to\lesssim1.75 microns (order 37 for GIANO-B) to repeat the CCF calculations. As shown in Figure 8, the result shows a lack of prominent signals at the expected locations when the wavelength coverage is insufficient.

Refer to caption
Figure 8: KpΔ\varvsubscript𝐾pΔ\varvK_{\mathrm{p}}-\Delta\varvitalic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT - roman_Δ maps of H2O for GIANO-B data (less-than-or-similar-to\lesssim1.75 microns) with two nights combined. The crossed white dashed lines indicate the S/N peak. The crossed black dashed lines indicate the expected Kpsubscript𝐾pK_{\mathrm{p}}italic_K start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT and Δ\varvΔ\varv\Delta\varvroman_Δ.

Appendix E Simulation of High-resolution Phase-resolved Transmission spectra with nightside thermal Emission

We conducted a phase-resolved high-resolution simulation of H2O thermal emission spectra by post-processing the GCM outputs with PICASO (Batalha et al., 2019; Mukherjee et al., 2023). We initially tested a reference GCM from Tan et al. (2024), including drag and H2-H dissociation and recombination, with an equilibrium temperature of 2600 K, close to that of WASP-33b, and a host star temperature of 6500 K (lower than WASP-33). The reference models include TiO and VO, indicating that the nightside upper-layer cooling is effective. The other parameter settings are described in Tan et al. (2024). We implemented radiative transfer calculations with 3D velocity fields by adapting the PICASO code with the radiative transfer methodology of Toon et al. (1989). For chemistry, we assumed chemical equilibrium without clouds (noting that the nightside of UHJs is typically covered with various high-temperature condensation clouds, to be discussed in a future paper). The line-by-line opacity database construction method can be found in the PICASO document333https://natashabatalha.github.io/picaso/notebooks/10_CreatingOpacityDb.html. The rotation angle during the transit of WASP-33b is about 35.06 (i.e., the transit duration over the period). We calculated the phase-resolved thermal emission spectra within ±plus-or-minus\pm±17.5 of the mid-transit phase with a step of 3.5. The classical transmission spectrum was simply calculated by petitRADTRANS. We compared the strength of the classical transmission spectrum with the nightside emission spectra in Figure 9.

Refer to caption
Figure 9: An example of phase-resolved general H2O transmission spectra simulation using a reference GCM with TiO and VO (i.e., nightside cooling is effective). Top panel: The normalized phase-resolved nightside thermal emission spectra from --17.5 to +17.5 with a step of 3.5 (color-coded from blue to red). Bottom panel: Phase-resolved general transmission spectra with nightside thermal emission. The black line shows the classical transmission spectrum for H2O lines (i.e., without considering night emission).