atex style emulateapj v. 12/16/11 - arxiv

17
Draft version October 15, 2021 Preprint typeset using L A T E X style emulateapj v. 12/16/11 TEMPORAL VARIABILITY IN HOT JUPITER ATMOSPHERES Thaddeus D. Komacek 1 , Adam P. Showman 2,3 1 Department of the Geophysical Sciences, The University of Chicago, Chicago, IL, 60637 2 Department of Atmospheric and Oceanic Sciences, School of Physics, Peking University, Beijing, 100871, China 3 Lunar and Planetary Laboratory and Department of Planetary Sciences, University of Arizona, Tucson, AZ, 85721 [email protected] Draft version October 15, 2021 ABSTRACT Hot Jupiters receive intense incident stellar light on their daysides, which drives vigorous atmo- spheric circulation that attempts to erase their large dayside-to-nightside flux contrasts. Propagating waves and instabilities in hot Jupiter atmospheres can cause emergent properties of the atmosphere to be time-variable. In this work, we study such weather in hot Jupiter atmospheres using idealized cloud-free general circulation models with double-grey radiative transfer. We find that hot Jupiter atmospheres can be time-variable at the 0.1 - 1% level in globally averaged temperature and at the 1 - 10% level in globally averaged wind speeds. As a result, we find that observable quantities are also time variable: the secondary eclipse depth can be variable at the . 2% level, the phase curve am- plitude can change by . 1%, the phase curve offset can shift by . 5 , and terminator-averaged wind speeds can vary by . 2 km s -1 . Additionally, we calculate how the eastern and western limb-averaged wind speeds vary with incident stellar flux and the strength of an imposed drag that parameterizes Lorentz forces in partially ionized atmospheres. We find that the eastern limb is blueshifted in mod- els over a wide range of equilibrium temperature and drag strength, while the western limb is only redshifted if equilibrium temperatures are . 1500 K and drag is weak. In general, our results show that the amplitude of time-variability increases with increasing incident stellar flux and decreasing atmospheric drag strength. We show that this variability may be observationally detectable in the infrared through secondary eclipse observations with JWST, phase curve observations with future space telescopes (e.g., ARIEL), and/or Doppler wind speed measurements with high-resolution spec- trographs. Subject headings: hydrodynamics - methods: numerical - planets and satellites: gaseous planets - planets and satellites: atmospheres 1. INTRODUCTION Planetary atmospheres are dynamic environments in which short-timescale temporal variability (i.e., “weather”) is ubiquitous. Given that storms and other large-scale weather patterns are triggered by atmospheric instabilities (Holton & Hakim 2013), time-variability in- forms us of the types of dynamical instabilities operating in planetary atmospheres. All Solar System planets ex- perience some form of atmospheric variability (Sanchez- Lavega 2010), and so one naturally expects weather to occur in the atmospheres of exoplanets as well. In this work, we study time-variability in the atmospheres of close-in gas giant planets, or “hot Jupiters.” Hot Jupiters are the hottest class of exoplanets and have large atmo- spheric scale heights, so the observational signatures of time-variability are likely to be most apparent in this class of planet. As a result, hot Jupiters can serve as a laboratory in which to study variability using a combi- nation of observational and theoretical approaches. To date, time-variability has been observed in the at- mospheres of three hot Jupiters: HAT-P-7b, Kepler-76b, and WASP-12b. Armstrong et al. (2016) and Jackson et al. (2019) found that the Kepler optical light curves of HAT-P-7b and Kepler-76b show a significant varia- tion in the phase offset, which is the temporal shift of the maximum brightness of the phase curve from sec- ondary eclipse. Further, the phase offset of HAT-P-7b and Kepler-76b shift from negative to positive, that is the time of maximum brightness shifts from before secondary eclipse to after secondary eclipse. Over a longer temporal baseline, Bell et al. (2019) found that the Spitzer 3.6 μm phase curve offset of WASP-12b shifted from 32.6 ± 6.2 eastward of the substellar point in 2010 to 13.6 ± 3.8 westward of the substellar point in 2013. Armstrong et al. (2016) and Jackson et al. (2019) note that this reversal could result from a changing cloud pattern or a change in the direction of the equatorial jet in the atmo- spheres of these likely tidally-locked planets. Addition- ally, Kepler-76b was found to have time-variability in the phase curve amplitude, while HAT-P-7b and WASP- 12b did not have significant variability in amplitude. Al- though atmospheric circulation models of hot Jupiters that exclude magnetic effects have almost universally shown that the equatorial jet should be eastward (su- perrotating), it has been shown that magnetic effects due to the interaction between an intrinsic magnetic field and the thermally ionized atmosphere can reverse the equato- rial jet in a time-dependent manner (Rogers & Komacek 2014, Hindle et al. 2019). Further, magnetic effects have been shown to qualitatively explain the phase offset vari- ability of HAT-P-7b (Rogers 2017). Additionally, significant time-variability has been ob- served in Spitzer infrared secondary eclipse observations of the hot super-Earth 55 Cancri e (Demory et al. 2016, Tamburo et al. 2018). The mechanism inducing this vari- ability has not been identified, but it has been hypoth- arXiv:1910.09523v1 [astro-ph.EP] 21 Oct 2019

Upload: others

Post on 06-Jan-2022

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: ATEX style emulateapj v. 12/16/11 - arXiv

Draft version October 15, 2021Preprint typeset using LATEX style emulateapj v. 12/16/11

TEMPORAL VARIABILITY IN HOT JUPITER ATMOSPHERES

Thaddeus D. Komacek1, Adam P. Showman2,3

1Department of the Geophysical Sciences, The University of Chicago, Chicago, IL, 606372Department of Atmospheric and Oceanic Sciences, School of Physics, Peking University, Beijing, 100871, China

3Lunar and Planetary Laboratory and Department of Planetary Sciences, University of Arizona, Tucson, AZ, [email protected]

Draft version October 15, 2021

ABSTRACT

Hot Jupiters receive intense incident stellar light on their daysides, which drives vigorous atmo-spheric circulation that attempts to erase their large dayside-to-nightside flux contrasts. Propagatingwaves and instabilities in hot Jupiter atmospheres can cause emergent properties of the atmosphereto be time-variable. In this work, we study such weather in hot Jupiter atmospheres using idealizedcloud-free general circulation models with double-grey radiative transfer. We find that hot Jupiteratmospheres can be time-variable at the ∼ 0.1− 1% level in globally averaged temperature and at the∼ 1− 10% level in globally averaged wind speeds. As a result, we find that observable quantities arealso time variable: the secondary eclipse depth can be variable at the . 2% level, the phase curve am-plitude can change by . 1%, the phase curve offset can shift by . 5, and terminator-averaged windspeeds can vary by . 2 km s−1. Additionally, we calculate how the eastern and western limb-averagedwind speeds vary with incident stellar flux and the strength of an imposed drag that parameterizesLorentz forces in partially ionized atmospheres. We find that the eastern limb is blueshifted in mod-els over a wide range of equilibrium temperature and drag strength, while the western limb is onlyredshifted if equilibrium temperatures are . 1500 K and drag is weak. In general, our results showthat the amplitude of time-variability increases with increasing incident stellar flux and decreasingatmospheric drag strength. We show that this variability may be observationally detectable in theinfrared through secondary eclipse observations with JWST, phase curve observations with futurespace telescopes (e.g., ARIEL), and/or Doppler wind speed measurements with high-resolution spec-trographs.Subject headings: hydrodynamics - methods: numerical - planets and satellites: gaseous planets -

planets and satellites: atmospheres

1. INTRODUCTION

Planetary atmospheres are dynamic environmentsin which short-timescale temporal variability (i.e.,“weather”) is ubiquitous. Given that storms and otherlarge-scale weather patterns are triggered by atmosphericinstabilities (Holton & Hakim 2013), time-variability in-forms us of the types of dynamical instabilities operatingin planetary atmospheres. All Solar System planets ex-perience some form of atmospheric variability (Sanchez-Lavega 2010), and so one naturally expects weather tooccur in the atmospheres of exoplanets as well. In thiswork, we study time-variability in the atmospheres ofclose-in gas giant planets, or “hot Jupiters.” Hot Jupitersare the hottest class of exoplanets and have large atmo-spheric scale heights, so the observational signatures oftime-variability are likely to be most apparent in thisclass of planet. As a result, hot Jupiters can serve as alaboratory in which to study variability using a combi-nation of observational and theoretical approaches.

To date, time-variability has been observed in the at-mospheres of three hot Jupiters: HAT-P-7b, Kepler-76b,and WASP-12b. Armstrong et al. (2016) and Jacksonet al. (2019) found that the Kepler optical light curvesof HAT-P-7b and Kepler-76b show a significant varia-tion in the phase offset, which is the temporal shift ofthe maximum brightness of the phase curve from sec-ondary eclipse. Further, the phase offset of HAT-P-7band Kepler-76b shift from negative to positive, that is the

time of maximum brightness shifts from before secondaryeclipse to after secondary eclipse. Over a longer temporalbaseline, Bell et al. (2019) found that the Spitzer 3.6 µmphase curve offset of WASP-12b shifted from 32.6± 6.2

eastward of the substellar point in 2010 to 13.6 ± 3.8

westward of the substellar point in 2013. Armstronget al. (2016) and Jackson et al. (2019) note that thisreversal could result from a changing cloud pattern or achange in the direction of the equatorial jet in the atmo-spheres of these likely tidally-locked planets. Addition-ally, Kepler-76b was found to have time-variability inthe phase curve amplitude, while HAT-P-7b and WASP-12b did not have significant variability in amplitude. Al-though atmospheric circulation models of hot Jupitersthat exclude magnetic effects have almost universallyshown that the equatorial jet should be eastward (su-perrotating), it has been shown that magnetic effects dueto the interaction between an intrinsic magnetic field andthe thermally ionized atmosphere can reverse the equato-rial jet in a time-dependent manner (Rogers & Komacek2014, Hindle et al. 2019). Further, magnetic effects havebeen shown to qualitatively explain the phase offset vari-ability of HAT-P-7b (Rogers 2017).

Additionally, significant time-variability has been ob-served in Spitzer infrared secondary eclipse observationsof the hot super-Earth 55 Cancri e (Demory et al. 2016,Tamburo et al. 2018). The mechanism inducing this vari-ability has not been identified, but it has been hypoth-

arX

iv:1

910.

0952

3v1

[as

tro-

ph.E

P] 2

1 O

ct 2

019

Page 2: ATEX style emulateapj v. 12/16/11 - arXiv

2 Komacek & Showman

esized by Demory et al. to be due to volcanic plumeserupting from the molten lithosphere of the planet.

To date, there has been no detection of time-variabilityin the secondary eclipse depth of hot Jupiters in theinfrared. Using a sample of seven Spitzer secondaryeclipses, Agol et al. (2010) placed an upper limit of2.7% on the variability in the secondary eclipse depthof HD 189733b at 8 µm. Kilpatrick et al. (2019) alsoused Spitzer observations to place an upper limit on thevariability of HD 189733b of 5.6% at 3.5µm and 6.0%at 4.5µm, and an upper limit on the variability of HD209458b of 12% at 3.5µm and 1.6% at 4.5µm. Thisvariability is consistent with that calculated in the gen-eral circulation modeling work of Showman et al. (2009),which predicts small variations in the secondary eclipsedepth of . 1%.

There are many dynamical processes at play that in-duce atmospheric variability in hot Jupiter atmospheres.These processes can be broken up into two main cat-egories: magnetohydrodynamic variability (due to thecoupling of magnetism and atmospheric dynamics) andhydrodynamic variability (due to purely dynamical pro-cesses). As discussed above, magnetic effects can in-duce time-variability by reversing the direction of theequatorial jet. On the hottest hot Jupiters, strong elec-trical currents can lead to the formation of an atmo-spheric dynamo, which would cause significant time-variability (Rogers & Mcelwaine 2017). Additionally,if the internally-generated dynamos of hot Jupiters aretilted with respect to the rotation axis, the flow is nolonger axisymmetric, leading to significant variability inthe latitudinal position of the equatorial jet (Batygin &Stanley 2014).

A wide range of purely hydrodynamic instabilities maybe at play in the atmospheres of hot Jupiters. First,horizontal shear instabilities can occur due to the largedifference in wind speed between the equatorial jet andits surroundings and induce barotropic eddies that sapenergy from the equatorial jet (Menou & Rauscher 2009,Fromang et al. 2016). Second, vertical shear instabilitiesin these stably stratified atmospheres may sap energyfrom the equatorial jet and limit its speed (Showman &Guillot 2002). Such vertical shear instabilities may causesignificant energy dissipation at low pressures, resultingin variability at potentially observable levels (Cho et al.2015, Fromang et al. 2016). Third, baroclinic instabilitieslikely play a role in driving jets in the atmospheres of gasgiant planets in the Solar System (Williams 1979, Kaspi& Flierl 2006, Lian & Showman 2008, Young et al. 2019),and could cause time-variability in the atmospheres ofhot Jupiters (Polichtchouk & Cho 2012). Fourth, thecloud pattern in hot Jupiter atmospheres may be patchy,and changes in this cloud pattern could lead to time-variability (Parmentier et al. 2013, Armstrong et al. 2016,Parmentier et al. 2016, Lines et al. 2018, Powell et al.2018). Lastly, the day-night forcing could be sufficientlyhigh amplitude such that the large-scale standing wavepattern that forces the equatorial jet becomes unstable(Showman & Polvani 2010). Note that even the simu-lations of Liu & Showman (2013) that did not includemagnetic effects or vertical shear instabilities found vari-ations in wind speeds on the order of a few percent.

With the high precision of the James Webb Space Tele-scope (JWST ) (Greene et al. 2016), we expect that sec-

ondary eclipse variability at the level of ≈ 1% will be ob-servable. If variability is detected, then the period andamplitude (or amplitude spectrum) of this variability canprovide critical information not obtainable in any otherway. In particular, the instability period and amplitudemay be influenced by basic-state atmospheric properties,such as the structure of the equatorial jet and verticalstratification. As a result, measurement of the variabil-ity properties may provide a constraint on basic-stateproperties of the atmosphere.

In this paper, we place constraints on the level of ob-servable time-variability in hot Jupiter atmospheres. Todo so, we analyze the time-variability of a large suiteof purely hydrodynamic simulations of the atmosphericcirculation of hot Jupiters. These simulations do resolvehorizontal shear instabilities, but because they are hydro-static cannot simulate vertical shear instabilities. Addi-tionally, the simulations do not include magnetic effectsand clouds, which could both lead to strong variabilityin the emitted planetary infrared flux. We also do notinclude the effects of hydrogen dissociation and recom-bination, which may cause emergent time-variability inultra-hot Jupiter atmospheres (Tan & Komacek 2019).As a result, one can consider our results as a baseline forunderstanding variability in a cloud-free and locally hy-drostatic atmosphere without the effects of magnetism.This work therefore provides a crucial point of compari-son for future models that study time-variability includ-ing clouds, non-hydrostatic dynamics, hydrogen dissoci-ation and recombination, and/or magnetic effects.

This paper is organized as follows. In Section 2, wedescribe our numerical setup and how we diagnose time-variability in our simulations. We then go on to analyzethe time-variability from our simulations, separating thistime-variability into two parts. First, we analyze thephysical (or “intrinsic”) time-variability in temperatureand wind speeds in Section 3.1. Then, in Section 3.2we show the observational consequences of this intrinsicvariability. We discuss the potential for observing time-variability with JWST and ground-based high resolutionspectroscopy, describe other mechanisms that may in-duce variability, and show model predictions for Dopplerwind speeds as a function of equilibrium temperature anddrag strength in Section 4. Lastly, we state key takeawaypoints in Section 5.

2. NUMERICAL SIMULATIONS

2.1. Numerical model

To model the time-variability in hot Jupiter atmo-spheres, we numerically solve the primitive equations ofmeteorology and represent the heating and cooling us-ing a double-grey two-stream radiative transfer scheme.Specifically, we use the same adapted version of thethree-dimensional General Circulation Model (GCM),the MITgcm (Adcroft et al. 2004), as is used in Komaceket al. (2017), Koll & Komacek (2018), and Komaceket al. (2019). This numerical model utilizes the dynami-cal core of the MITgcm to solve the hydrostatic primitiveequations of atmospheric motion on a cubed-sphere grid.Additionally, we use the TWOSTR radiative transfer pack-age (Kylling et al. 1995) of the DISORT radiative transfercode (Stamnes et al. 1988) to solve the plane-parallel ra-diative transfer equations with the two-stream approxi-

Page 3: ATEX style emulateapj v. 12/16/11 - arXiv

Time Variability in Hot Jupiter Atmospheres 3

mation. We further use the double-grey approximation,considering only two radiative bands, one for the visibleand one for the infrared. This approximation has beenused previously for studies of hot Jupiter atmospheres(Heng et al. 2011, Rauscher & Menou 2012, Komaceket al. 2017, Tan & Komacek 2019), and we use the sameparameterizations for the visible and infrared absorptioncoefficients as in Komacek et al. (2017).

2.2. Physical parameters

We use the same planetary parameters relevant forHD 209458b as in Komacek & Showman (2016) andKomacek et al. (2017) for all simulations. We usea rotation rate of Ω = 2.078 × 10−5 s−1, grav-ity of g = 9.36 m s−2, and planetary radius of a =9.437 × 107 m. We further vary the irradiation cor-responding to model equilibrium temperatures of Teq =500, 1000, 1500, 2000, 2500, and 3000 K, equivalent tovarying the incident stellar flux from 1.42× 104− 1.84×107 W m−2. We use the same absorption coefficients asin Komacek et al. (2017), chosen in order to match theresults of more detailed radiative transfer calculations.The absorption coefficient for incoming visible radiationis fixed at κv = 4 × 10−4 m2kg−1. The absorption co-efficient for outgoing thermal radiation is set to a powerlaw in pressure, κth = 2.28× 10−6 (p/1 Pa)

0.53m2kg−1.

These absorption coefficients correspond to a visible pho-tosphere that lies at a pressure of 0.23 bars and a ther-mal photosphere at 0.21 bars. We include a small butnon-negligible internal heat flux in our radiative transfercalculation. Our chosen internal heat flux is equivalentto an internal effective temperature of Tint = 100 K, sim-ilar to that of Jupiter. As a result, we do not include theenhancement in the internal heat flux of hot Jupiters dueto internal heat deposition (Thorngren et al. 2019).

We also impose a Rayleigh drag force that is linearin wind speed and has a magnitude inversely propor-tional to an assumed drag timescale τdrag. As in Ko-macek et al. (2017), we consider a range of τdrag =103, 104, 105, 106, 107, and∞ s. If τdrag ≤ 105 s, the dragtimescale is spatially uniform. However, in simulationswith τdrag > 105 s, we also include a basal drag term.This basal drag was first introduced by Liu & Showman(2013) in order to ensure that the equilibrated end-stateof hot Jupiter simulations are independent of initial con-ditions and to crudely parameterize the interactions be-tween the deep circulation and the interior of the planet.In relaxing the deep winds (near the base of the model)toward zero, we are essentially presuming that the inte-rior has weak winds due to the weak internal heat fluxesand significant expected drag due to Lorentz forces in theconvective region (e.g., Liu et al. 2008). As in Komacek& Showman (2016), we assume that this basal drag in-creases in strength as a power-law from no drag at apressure of 10 bars to τdrag,basal = 10 days at the bottomof the domain, which occurs at a pressure of 200 bars.Note that this basal drag term acts to damp atmosphericvariability deep in the atmosphere (at p ≥ 10 bars). Weran test simulations with τdrag,basal = 100 days and foundthat the variability pattern and amplitude near the pho-tosphere is not greatly affected by reducing the strengthof basal drag. Our applied Rayleigh drag timescale isconstant with time throughout all simulations.

2.3. Numerical parameters

In our suite of numerical experiments, we use a numer-ical resolution of C32, which corresponds to 32×32 finitevolume elements on each cube face of the cubed spheregrid and approximately equals a resolution of 128 longi-tudinal and 64 latitudinal grid cells. We use 40 verticallevels, 39 of which are evenly spaced in log-pressure from200 bars to 0.2 mbars and with a top layer extendingto zero pressure. These simulations include a fourth-order Shapiro filter (Shapiro 1971) to damp grid-scalevariations. In Koll & Komacek (2018) it was shown thatthis Shapiro filter does not affect the global angular mo-mentum budget, but can affect kinetic energy dissipationrates, especially for simulations with τdrag ≥ 105 s. Asa result, we computed test cases with an increased hori-zontal resolution of C64 and found that the amplitude oftemperature and wind speed variability calculated fromour nominal simulations with a horizontal resolution ofC32 agree with those using a higher resolution. We runour simulations for a model time of 2,000 Earth days tospin-up the model to reach an equilibrated state in ki-netic energy at observable levels (see Section 2.4 for moredetails). To calculate the resulting time-variability fromour simulations, we output model results (e.g., temper-ature and wind speeds) every 12 hours for the last 150days of the simulation.

2.4. Checking energy equilibration

As in Liu & Showman (2013), we check the time-evolution of our simulations to determine if kinetic en-ergy is equilibrated. We calculate the domain-integratedkinetic energy1, KEtot, as

KEtot =

p

A

u2 + v2

2gdA dp, (1)

where u and v are the zonal (east-west) and merid-ional (north-south) wind speeds at each grid point, g =9.36 m s−2 is the assumed surface gravity, A is horizon-tal area, and p is pressure. All simulations in the suiteof models reach a steady-state in domain-integrated ki-netic energy by 2,000 days of model time, except weaklyforced simulations with Teq = 500 K and weakly dampedsimulations with τdrag =∞. As a result, we do not showresults from the weakly forced suite of simulations withTeq = 500 K. Note that hot Jupiters are defined to haveTeq ≥ 1000 K, so the simulations with Teq = 500 K rep-resent tidally-locked gas giant planets that receive signif-icantly less incident stellar radiation than hot Jupiters.

3. RESULTS

3.1. Intrinsic time-variability

3.1.1. Temperature pattern

Now we turn to analyze the intrinsic variability in hotJupiter atmospheres, first studying the time-variabilityin the temperature pattern of hot Jupiters. Figure 1(top panel) shows maps of the time-variability in tem-perature at a pressure of 80 mbar for a simulation with

1 In the primitive equations, energetic self-consistency impliesthat the proper expression for local kinetic energy is KE =(u2 + v2

)/2 and hence does not include a term for vertical ve-

locity (Holton & Hakim 2013).

Page 4: ATEX style emulateapj v. 12/16/11 - arXiv

4 Komacek & Showman

T [K]t = tfin t = tfin 150 days t = tfin 75 days(a) (b) (c)

Longitude (°)

Latit

ude

(° )

−180 −90 0 90 180−80

−60

−40

−20

0

20

40

60

80

1100

1200

1300

1400

1500

1600

1700

1800

Longitude (°)

Latit

ude

(° )

−180 −90 0 90 180−80

−60

−40

−20

0

20

40

60

80

−2

−1.5

−1

−0.5

0

0.5

1

1.5

2

Longitude (°)

Latit

ude

(° )

−180 −90 0 90 180−80

−60

−40

−20

0

20

40

60

80

−30

−20

−10

0

10

20

30

Longitude (°)

Latit

ude

(° )

−180 −90 0 90 180−80

−60

−40

−20

0

20

40

60

80

−30

−20

−10

0

10

20

30T [K]

(d)<latexit sha1_base64="9/l0qNLfXJGvkg/w/SCKJYiYQ7Q=">AAAB8nicbVBNSwMxFHzrZ61fVY9egkWol7IVQb0VvXis4NpCdynZbLYNTbJLkhXK0r/hxYOKV3+NN/+N2XYP2joQGGbe400mTDnTxnW/nZXVtfWNzcpWdXtnd2+/dnD4qJNMEeqRhCeqF2JNOZPUM8xw2ksVxSLktBuObwu/+0SVZol8MJOUBgIPJYsZwcZKfsMX2IyUyKPp2aBWd5vuDGiZtEpShxKdQe3LjxKSCSoN4VjrfstNTZBjZRjhdFr1M01TTMZ4SPuWSiyoDvJZ5ik6tUqE4kTZJw2aqb83ciy0nojQThYR9aJXiP95/czEV0HOZJoZKsn8UJxxZBJUFIAipigxfGIJJorZrIiMsMLE2JqqtoTW4peXiXfevG669xf19k3ZRgWO4QQa0IJLaMMddMADAik8wyu8OZnz4rw7H/PRFafcOYI/cD5/ADavkVc=</latexit><latexit sha1_base64="9/l0qNLfXJGvkg/w/SCKJYiYQ7Q=">AAAB8nicbVBNSwMxFHzrZ61fVY9egkWol7IVQb0VvXis4NpCdynZbLYNTbJLkhXK0r/hxYOKV3+NN/+N2XYP2joQGGbe400mTDnTxnW/nZXVtfWNzcpWdXtnd2+/dnD4qJNMEeqRhCeqF2JNOZPUM8xw2ksVxSLktBuObwu/+0SVZol8MJOUBgIPJYsZwcZKfsMX2IyUyKPp2aBWd5vuDGiZtEpShxKdQe3LjxKSCSoN4VjrfstNTZBjZRjhdFr1M01TTMZ4SPuWSiyoDvJZ5ik6tUqE4kTZJw2aqb83ciy0nojQThYR9aJXiP95/czEV0HOZJoZKsn8UJxxZBJUFIAipigxfGIJJorZrIiMsMLE2JqqtoTW4peXiXfevG669xf19k3ZRgWO4QQa0IJLaMMddMADAik8wyu8OZnz4rw7H/PRFafcOYI/cD5/ADavkVc=</latexit><latexit sha1_base64="9/l0qNLfXJGvkg/w/SCKJYiYQ7Q=">AAAB8nicbVBNSwMxFHzrZ61fVY9egkWol7IVQb0VvXis4NpCdynZbLYNTbJLkhXK0r/hxYOKV3+NN/+N2XYP2joQGGbe400mTDnTxnW/nZXVtfWNzcpWdXtnd2+/dnD4qJNMEeqRhCeqF2JNOZPUM8xw2ksVxSLktBuObwu/+0SVZol8MJOUBgIPJYsZwcZKfsMX2IyUyKPp2aBWd5vuDGiZtEpShxKdQe3LjxKSCSoN4VjrfstNTZBjZRjhdFr1M01TTMZ4SPuWSiyoDvJZ5ik6tUqE4kTZJw2aqb83ciy0nojQThYR9aJXiP95/czEV0HOZJoZKsn8UJxxZBJUFIAipigxfGIJJorZrIiMsMLE2JqqtoTW4peXiXfevG669xf19k3ZRgWO4QQa0IJLaMMddMADAik8wyu8OZnz4rw7H/PRFafcOYI/cD5/ADavkVc=</latexit>

(e)<latexit sha1_base64="EpbCsYS7ROA4tkkCEDnMf7n3gb8=">AAAB8nicbVBNSwMxFHzrZ61fVY9egkWol7IrgnorevFYwbWF7lKyabYNTbJLkhXK0r/hxYOKV3+NN/+N2XYP2joQGGbe400mSjnTxnW/nZXVtfWNzcpWdXtnd2+/dnD4qJNMEeqThCeqG2FNOZPUN8xw2k0VxSLitBONbwu/80SVZol8MJOUhgIPJYsZwcZKQSMQ2IyUyOn0rF+ru013BrRMvJLUoUS7X/sKBgnJBJWGcKx1z3NTE+ZYGUY4nVaDTNMUkzEe0p6lEguqw3yWeYpOrTJAcaLskwbN1N8bORZaT0RkJ4uIetErxP+8XmbiqzBnMs0MlWR+KM44MgkqCkADpigxfGIJJorZrIiMsMLE2JqqtgRv8cvLxD9vXjfd+4t666ZsowLHcAIN8OASWnAHbfCBQArP8ApvTua8OO/Ox3x0xSl3juAPnM8fODSRWA==</latexit><latexit sha1_base64="EpbCsYS7ROA4tkkCEDnMf7n3gb8=">AAAB8nicbVBNSwMxFHzrZ61fVY9egkWol7IrgnorevFYwbWF7lKyabYNTbJLkhXK0r/hxYOKV3+NN/+N2XYP2joQGGbe400mSjnTxnW/nZXVtfWNzcpWdXtnd2+/dnD4qJNMEeqThCeqG2FNOZPUN8xw2k0VxSLitBONbwu/80SVZol8MJOUhgIPJYsZwcZKQSMQ2IyUyOn0rF+ru013BrRMvJLUoUS7X/sKBgnJBJWGcKx1z3NTE+ZYGUY4nVaDTNMUkzEe0p6lEguqw3yWeYpOrTJAcaLskwbN1N8bORZaT0RkJ4uIetErxP+8XmbiqzBnMs0MlWR+KM44MgkqCkADpigxfGIJJorZrIiMsMLE2JqqtgRv8cvLxD9vXjfd+4t666ZsowLHcAIN8OASWnAHbfCBQArP8ApvTua8OO/Ox3x0xSl3juAPnM8fODSRWA==</latexit><latexit sha1_base64="EpbCsYS7ROA4tkkCEDnMf7n3gb8=">AAAB8nicbVBNSwMxFHzrZ61fVY9egkWol7IrgnorevFYwbWF7lKyabYNTbJLkhXK0r/hxYOKV3+NN/+N2XYP2joQGGbe400mSjnTxnW/nZXVtfWNzcpWdXtnd2+/dnD4qJNMEeqThCeqG2FNOZPUN8xw2k0VxSLitBONbwu/80SVZol8MJOUhgIPJYsZwcZKQSMQ2IyUyOn0rF+ru013BrRMvJLUoUS7X/sKBgnJBJWGcKx1z3NTE+ZYGUY4nVaDTNMUkzEe0p6lEguqw3yWeYpOrTJAcaLskwbN1N8bORZaT0RkJ4uIetErxP+8XmbiqzBnMs0MlWR+KM44MgkqCkADpigxfGIJJorZrIiMsMLE2JqqtgRv8cvLxD9vXjfd+4t666ZsowLHcAIN8OASWnAHbfCBQArP8ApvTua8OO/Ox3x0xSl3juAPnM8fODSRWA==</latexit>

(f)<latexit sha1_base64="FieO8EaKFNTbbRG1NKTVggvL684=">AAAB8nicbVBNSwMxFHzrZ61fVY9egkWol7IrgnorevFYwbWF7lKyabYNTbJLkhXK0r/hxYOKV3+NN/+N2XYP2joQGGbe400mSjnTxnW/nZXVtfWNzcpWdXtnd2+/dnD4qJNMEeqThCeqG2FNOZPUN8xw2k0VxSLitBONbwu/80SVZol8MJOUhgIPJYsZwcZKQSMQ2IyUyOPpWb9Wd5vuDGiZeCWpQ4l2v/YVDBKSCSoN4VjrnuemJsyxMoxwOq0GmaYpJmM8pD1LJRZUh/ks8xSdWmWA4kTZJw2aqb83ciy0nojIThYR9aJXiP95vczEV2HOZJoZKsn8UJxxZBJUFIAGTFFi+MQSTBSzWREZYYWJsTVVbQne4peXiX/evG669xf11k3ZRgWO4QQa4MEltOAO2uADgRSe4RXenMx5cd6dj/noilPuHMEfOJ8/ObmRWQ==</latexit><latexit sha1_base64="FieO8EaKFNTbbRG1NKTVggvL684=">AAAB8nicbVBNSwMxFHzrZ61fVY9egkWol7IrgnorevFYwbWF7lKyabYNTbJLkhXK0r/hxYOKV3+NN/+N2XYP2joQGGbe400mSjnTxnW/nZXVtfWNzcpWdXtnd2+/dnD4qJNMEeqThCeqG2FNOZPUN8xw2k0VxSLitBONbwu/80SVZol8MJOUhgIPJYsZwcZKQSMQ2IyUyOPpWb9Wd5vuDGiZeCWpQ4l2v/YVDBKSCSoN4VjrnuemJsyxMoxwOq0GmaYpJmM8pD1LJRZUh/ks8xSdWmWA4kTZJw2aqb83ciy0nojIThYR9aJXiP95vczEV2HOZJoZKsn8UJxxZBJUFIAGTFFi+MQSTBSzWREZYYWJsTVVbQne4peXiX/evG669xf11k3ZRgWO4QQa4MEltOAO2uADgRSe4RXenMx5cd6dj/noilPuHMEfOJ8/ObmRWQ==</latexit><latexit sha1_base64="FieO8EaKFNTbbRG1NKTVggvL684=">AAAB8nicbVBNSwMxFHzrZ61fVY9egkWol7IrgnorevFYwbWF7lKyabYNTbJLkhXK0r/hxYOKV3+NN/+N2XYP2joQGGbe400mSjnTxnW/nZXVtfWNzcpWdXtnd2+/dnD4qJNMEeqThCeqG2FNOZPUN8xw2k0VxSLitBONbwu/80SVZol8MJOUhgIPJYsZwcZKQSMQ2IyUyOPpWb9Wd5vuDGiZeCWpQ4l2v/YVDBKSCSoN4VjrnuemJsyxMoxwOq0GmaYpJmM8pD1LJRZUh/ks8xSdWmWA4kTZJw2aqb83ciy0nojIThYR9aJXiP95vczEV2HOZJoZKsn8UJxxZBJUFIAGTFFi+MQSTBSzWREZYYWJsTVVbQne4peXiX/evG669xf11k3ZRgWO4QQa4MEltOAO2uADgRSe4RXenMx5cd6dj/noilPuHMEfOJ8/ObmRWQ==</latexit>

U [ms1]<latexit sha1_base64="jSh2aNBrRqlOCR+xtO7QBhy6e2E=">AAACBnicbZBPS8MwGMZT/875r+pRhOAQvDhaEfQ49OJxgt0GXR1plm5hSVqSVBilJy9+FS8eFPHqZ/DmtzHdKujmA4Efz/u+5H2fMGFUacf5shYWl5ZXVitr1fWNza1te2e3peJUYuLhmMWyEyJFGBXE01Qz0kkkQTxkpB2Orop6+55IRWNxq8cJCTgaCBpRjLSxevaBB/0uR3ooecbzH1L5XXbi5kHPrjl1ZyI4D24JNVCq2bM/u/0Yp5wIjRlSynedRAcZkppiRvJqN1UkQXiEBsQ3KBAnKsgmZ+TwyDh9GMXSPKHhxP09kSGu1JiHprPYU83WCvO/mp/q6CLIqEhSTQSefhSlDOoYFpnAPpUEazY2gLCkZleIh0girE1yVROCO3vyPLRO665Td2/Oao3LMo4K2AeH4Bi44Bw0wDVoAg9g8ACewAt4tR6tZ+vNep+2LljlzB74I+vjG7vsmUg=</latexit><latexit sha1_base64="jSh2aNBrRqlOCR+xtO7QBhy6e2E=">AAACBnicbZBPS8MwGMZT/875r+pRhOAQvDhaEfQ49OJxgt0GXR1plm5hSVqSVBilJy9+FS8eFPHqZ/DmtzHdKujmA4Efz/u+5H2fMGFUacf5shYWl5ZXVitr1fWNza1te2e3peJUYuLhmMWyEyJFGBXE01Qz0kkkQTxkpB2Orop6+55IRWNxq8cJCTgaCBpRjLSxevaBB/0uR3ooecbzH1L5XXbi5kHPrjl1ZyI4D24JNVCq2bM/u/0Yp5wIjRlSynedRAcZkppiRvJqN1UkQXiEBsQ3KBAnKsgmZ+TwyDh9GMXSPKHhxP09kSGu1JiHprPYU83WCvO/mp/q6CLIqEhSTQSefhSlDOoYFpnAPpUEazY2gLCkZleIh0girE1yVROCO3vyPLRO665Td2/Oao3LMo4K2AeH4Bi44Bw0wDVoAg9g8ACewAt4tR6tZ+vNep+2LljlzB74I+vjG7vsmUg=</latexit><latexit sha1_base64="jSh2aNBrRqlOCR+xtO7QBhy6e2E=">AAACBnicbZBPS8MwGMZT/875r+pRhOAQvDhaEfQ49OJxgt0GXR1plm5hSVqSVBilJy9+FS8eFPHqZ/DmtzHdKujmA4Efz/u+5H2fMGFUacf5shYWl5ZXVitr1fWNza1te2e3peJUYuLhmMWyEyJFGBXE01Qz0kkkQTxkpB2Orop6+55IRWNxq8cJCTgaCBpRjLSxevaBB/0uR3ooecbzH1L5XXbi5kHPrjl1ZyI4D24JNVCq2bM/u/0Yp5wIjRlSynedRAcZkppiRvJqN1UkQXiEBsQ3KBAnKsgmZ+TwyDh9GMXSPKHhxP09kSGu1JiHprPYU83WCvO/mp/q6CLIqEhSTQSefhSlDOoYFpnAPpUEazY2gLCkZleIh0girE1yVROCO3vyPLRO665Td2/Oao3LMo4K2AeH4Bi44Bw0wDVoAg9g8ACewAt4tR6tZ+vNep+2LljlzB74I+vjG7vsmUg=</latexit><latexit sha1_base64="jSh2aNBrRqlOCR+xtO7QBhy6e2E=">AAACBnicbZBPS8MwGMZT/875r+pRhOAQvDhaEfQ49OJxgt0GXR1plm5hSVqSVBilJy9+FS8eFPHqZ/DmtzHdKujmA4Efz/u+5H2fMGFUacf5shYWl5ZXVitr1fWNza1te2e3peJUYuLhmMWyEyJFGBXE01Qz0kkkQTxkpB2Orop6+55IRWNxq8cJCTgaCBpRjLSxevaBB/0uR3ooecbzH1L5XXbi5kHPrjl1ZyI4D24JNVCq2bM/u/0Yp5wIjRlSynedRAcZkppiRvJqN1UkQXiEBsQ3KBAnKsgmZ+TwyDh9GMXSPKHhxP09kSGu1JiHprPYU83WCvO/mp/q6CLIqEhSTQSefhSlDOoYFpnAPpUEazY2gLCkZleIh0girE1yVROCO3vyPLRO665Td2/Oao3LMo4K2AeH4Bi44Bw0wDVoAg9g8ACewAt4tR6tZ+vNep+2LljlzB74I+vjG7vsmUg=</latexit>

U [ms1]<latexit sha1_base64="ZPS0DkKMxxDxxtmbrJhZpk9N8oc=">AAACDXicbZDNSsNAFIUn9a/Wv6hLN4NVcGNJRNBlURcuK5i2kMQymU7aoTNJmJkIJeQF3Pgqblwo4ta9O9/GSRtBWw8MfJx7L3PvCRJGpbKsL6OysLi0vFJdra2tb2xumds7bRmnAhMHxywW3QBJwmhEHEUVI91EEMQDRjrB6LKod+6JkDSObtU4IT5Hg4iGFCOlrZ554F0RphB0oOtxpIaCZzz/IZnfZcd27vfMutWwJoLzYJdQB6VaPfPT68c45SRSmCEpXdtKlJ8hoShmJK95qSQJwiM0IK7GCHEi/WxyTQ4PtdOHYSz0ixScuL8nMsSlHPNAdxZ7ytlaYf5Xc1MVnvsZjZJUkQhPPwpTBlUMi2hgnwqCFRtrQFhQvSvEQyQQVjrAmg7Bnj15HtonDdtq2Den9eZFGUcV7IF9cARscAaa4Bq0gAMweABP4AW8Go/Gs/FmvE9bK0Y5swv+yPj4BrDBm/Q=</latexit><latexit sha1_base64="ZPS0DkKMxxDxxtmbrJhZpk9N8oc=">AAACDXicbZDNSsNAFIUn9a/Wv6hLN4NVcGNJRNBlURcuK5i2kMQymU7aoTNJmJkIJeQF3Pgqblwo4ta9O9/GSRtBWw8MfJx7L3PvCRJGpbKsL6OysLi0vFJdra2tb2xumds7bRmnAhMHxywW3QBJwmhEHEUVI91EEMQDRjrB6LKod+6JkDSObtU4IT5Hg4iGFCOlrZ554F0RphB0oOtxpIaCZzz/IZnfZcd27vfMutWwJoLzYJdQB6VaPfPT68c45SRSmCEpXdtKlJ8hoShmJK95qSQJwiM0IK7GCHEi/WxyTQ4PtdOHYSz0ixScuL8nMsSlHPNAdxZ7ytlaYf5Xc1MVnvsZjZJUkQhPPwpTBlUMi2hgnwqCFRtrQFhQvSvEQyQQVjrAmg7Bnj15HtonDdtq2Den9eZFGUcV7IF9cARscAaa4Bq0gAMweABP4AW8Go/Gs/FmvE9bK0Y5swv+yPj4BrDBm/Q=</latexit><latexit sha1_base64="ZPS0DkKMxxDxxtmbrJhZpk9N8oc=">AAACDXicbZDNSsNAFIUn9a/Wv6hLN4NVcGNJRNBlURcuK5i2kMQymU7aoTNJmJkIJeQF3Pgqblwo4ta9O9/GSRtBWw8MfJx7L3PvCRJGpbKsL6OysLi0vFJdra2tb2xumds7bRmnAhMHxywW3QBJwmhEHEUVI91EEMQDRjrB6LKod+6JkDSObtU4IT5Hg4iGFCOlrZ554F0RphB0oOtxpIaCZzz/IZnfZcd27vfMutWwJoLzYJdQB6VaPfPT68c45SRSmCEpXdtKlJ8hoShmJK95qSQJwiM0IK7GCHEi/WxyTQ4PtdOHYSz0ixScuL8nMsSlHPNAdxZ7ytlaYf5Xc1MVnvsZjZJUkQhPPwpTBlUMi2hgnwqCFRtrQFhQvSvEQyQQVjrAmg7Bnj15HtonDdtq2Den9eZFGUcV7IF9cARscAaa4Bq0gAMweABP4AW8Go/Gs/FmvE9bK0Y5swv+yPj4BrDBm/Q=</latexit><latexit sha1_base64="ZPS0DkKMxxDxxtmbrJhZpk9N8oc=">AAACDXicbZDNSsNAFIUn9a/Wv6hLN4NVcGNJRNBlURcuK5i2kMQymU7aoTNJmJkIJeQF3Pgqblwo4ta9O9/GSRtBWw8MfJx7L3PvCRJGpbKsL6OysLi0vFJdra2tb2xumds7bRmnAhMHxywW3QBJwmhEHEUVI91EEMQDRjrB6LKod+6JkDSObtU4IT5Hg4iGFCOlrZ554F0RphB0oOtxpIaCZzz/IZnfZcd27vfMutWwJoLzYJdQB6VaPfPT68c45SRSmCEpXdtKlJ8hoShmJK95qSQJwiM0IK7GCHEi/WxyTQ4PtdOHYSz0ixScuL8nMsSlHPNAdxZ7ytlaYf5Xc1MVnvsZjZJUkQhPPwpTBlUMi2hgnwqCFRtrQFhQvSvEQyQQVjrAmg7Bnj15HtonDdtq2Den9eZFGUcV7IF9cARscAaa4Bq0gAMweABP4AW8Go/Gs/FmvE9bK0Y5swv+yPj4BrDBm/Q=</latexit>

Longitude (°)

Latit

ude

(° )

−180 −90 0 90 180−80

−60

−40

−20

0

20

40

60

80

0

1000

2000

3000

4000

Longitude (°)

Latit

ude

(° )

−180 −90 0 90 180−80

−60

−40

−20

0

20

40

60

80

−100

−80

−60

−40

−20

0

20

40

60

Longitude (°)

Latit

ude

(° )

−180 −90 0 90 180−80

−60

−40

−20

0

20

40

60

80

−100

−80

−60

−40

−20

0

20

40

60

Fig. 1.— Maps showing the typical temperature pattern and the characteristic time-variation in temperature (top) along with the typicalzonal wind pattern and the characteristic time-variation in zonal wind (bottom) at 80 mbar pressure for the simulation with Teq = 1500 Kand τdrag = ∞ (i.e. no applied drag in the free atmosphere). (a) Temperature map with over-plotted wind vectors at the end of thesimulation, t = tfin = 2000 days. The characteristic superrotating jet and eastward hot spot offset of hot Jupiters is apparent. (b,c) Map ofthe absolute change in temperature between t = tfin − 150 days and t = tfin (b) and between t = tfin − 75 days and t = tfin (c). (d) Zonalwind map at the end of the simulation. (e,f) Same as (b,c) but showing maps of the absolute change in zonal wind speed. Temperatureand zonal wind changes are highest in the equatorial regions and occur at the 2% level or lower. Note that the change in temperature islargest near the western terminator, where there is strong convergence because the equatorial jet slows down as it passes the day-nightterminator.

drag = 1 drag = 106 s

Longitude (°)

Tim

e (d

ays)

−100 0 1001860

1880

1900

1920

1940

1960

1980

−500

−400

−300

−200

−100

0

100

200

300

Longitude (°)

Tim

e (d

ays)

−100 0 1001860

1880

1900

1920

1940

1960

1980

−500

−400

−300

−200

−100

0

100

200

300

T

T(t

)[K

]

Fig. 2.— Time-longitude section, also known as a Hovmoller diagram, of temperature deviations from the mean longitude-averagedtemperature at a given time at a pressure of 80 mbar, averaged over latitudes of ±15 from the equator. Shown here are results forsimulations with Teq = 1500 K and τdrag =∞ and 106 s. Variability for both simulations is large at the western terminator (at a longitude

of −90). The amplitude of variability is significantly larger in the case with τdrag = 106 s, partially resulting from the much largertemperature contrast at the western terminator in this simulation.

Page 5: ATEX style emulateapj v. 12/16/11 - arXiv

Time Variability in Hot Jupiter Atmospheres 5

1000 1500 2000 2500 30000.0

0.5

1.0

1.5

2.0

Glo

bal

∆T

rms/T

rms,

max

[%]

p = 1 bar

1000 1500 2000 2500 30000.0

0.5

1.0

1.5

2.0

2.5

p = 80 mbars

τdrag =∞τdrag = 107 s

τdrag = 106 s

τdrag = 105 s

τdrag = 104 s

τdrag = 103 s

1000 1500 2000 2500 30000

2

4

6p = 1 mbar

1000 1500 2000 2500 30000.0

0.5

1.0

1.5

2.0

Day

sid

e∆T

rms/T

rms,

max

[%]

1000 1500 2000 2500 30000.0

0.5

1.0

1.5

2.0

2.5

1000 1500 2000 2500 30000

2

4

6

1000 1500 2000 2500 30000.0

0.5

1.0

1.5

2.0

Nig

htsi

de

∆T

rms/T

rms,

max

[%]

1000 1500 2000 2500 30000.0

0.5

1.0

1.5

2.0

2.5

1000 1500 2000 2500 30000

2

4

6

1000 1500 2000 2500 3000Equilibrium Temperature (Kelvin)

0.0

0.5

1.0

1.5

2.0

Loc

alT

rms,

loc/T

av[%

]

1000 1500 2000 2500 3000Equilibrium Temperature (Kelvin)

0.0

0.5

1.0

1.5

2.0

2.5

1000 1500 2000 2500 3000Equilibrium Temperature (Kelvin)

0

2

4

6

Fig. 3.— Normalized amplitude of variability in the horizontal root-mean-square of temperature (top three rows) and locallized root-mean-square of temperature (bottom row) at given pressure levels over the last 150 days of simulation time. The temperature variability isshown for all for simulations with varying τdrag = 103−∞ s and Teq = 1000−3000 K at three different pressures: p = 1 bar, 80 mbars, and1 mbar. Plots in each column share a y-axis scale. The top row shows the global amplitude of variability, the second row shows the daysidevariability amplitude, the third row shows the nightside amplitude of variability, and the bottom row shows a horizontal root-mean-squareof the local measure of variability defined in Equation (4). The temperature variability generally increases with decreasing pressure level,and can be & 1% at p ≤ 1 mbar. At pressures ≥ 80 mbar, the variability generally increases with increasing equilibrium temperature andis largest for the case with τdrag = 106 s. At a pressure of 1 mbar and Teq ≥ 1500 K, the global variability is largest for the case without

applied drag in the free atmosphere (τdrag = ∞). Cases with strong drag (τdrag ≤ 105 s) generally show extremely little temperaturevariability. Small-scale variability (bottom row) is always present, even in simulations with strong drag.

Page 6: ATEX style emulateapj v. 12/16/11 - arXiv

6 Komacek & Showman

Teq = 1500 K and τdrag =∞, approximately representingthe actual conditions of HD 209458b. Locally, there is. 2% variability over 150 days of atmospheric evolutionin both temperature and zonal winds. This variability islargely confined to the equatorial regions. Notably, theequatorial regions are where the two most prominent fea-tures of hot Jupiter atmospheric circulation lie: a nearlysteady equatorial planetary-scale standing wave patternand a superrotating equatorial jet (Showman et al. 2010,Heng & Showman 2015). Additionally, note that the am-plitude of variability is largest in localized regions nearthe western and eastern terminators.

To study in more detail how the longitudinal tem-perature distribution changes with time, we producetime-longitude cross sections of temperature for two ofour simulations. Figure 2 shows these Hovmoller dia-grams of temperature at 80 mbar pressure over the last150 Earth days of simulation time for simulations withTeq = 1500 K and τdrag = ∞ and 106 s. In Figure 2,the temperature at any given longitude is averaged be-tween equatorial latitudes of ±15, where the variabilityis strongest. The temperature variability is significantlystronger for the case with τdrag = 106 s.

As discussed above, the variability (which is more vis-ible in the panel with τdrag = 106 s) is strongest at thewestern terminator (at a longitude of−90) and eastwardof the eastern terminator (at a longitude of ∼ 130). Thevariability at the western terminator occurs where thereis a sharp temperature boundary as air coming from thecool nightside hits the hot, irradiated dayside. The fea-ture past the eastern terminator is where air from higherlevels is advected downward. This downward verticaladvection results in a localized increase in the temper-ature relative to the cooler nightside surrounding air asa result of adiabatic heating. In general, the large-scaleequatorial standing wave pattern drives this circulation(Showman et al. 2008, Rauscher & Menou 2010, Show-man & Polvani 2011), which is known as a Gill patternin the Earth tropical dynamics literature (Matsuno 1966,Gill 1980).

This planetary-scale equatorial standing wave (“Gill”)pattern is analogous to those in Earth’s tropics in that itis comprised of equatorial Rossby and Kelvin wave modes(Showman & Polvani 2010, 2011, Tsai et al. 2014, Ham-mond & Pierrehumbert 2018). These waves are triggeredby the large day-to-night forcing contrast on the planet,and are damped by radiative cooling and frictional drag(Perez-Becker & Showman 2013, Komacek & Showman2016). Most of the time-variability in our simulations islikely due to changes in this large-scale standing wavepattern. Though in this work we do not develop a de-tailed theory for the variability, we speculate on possiblemechanisms in Section 4.1.

The time-variability in equatorial regions is significantenough that it can affect the globally averaged temper-ature. In Figure 3, we plot the variations in the root-mean-square (rms) temperature as a function of pressure,defined as

Trms(p) =

√∫T 2dA

A, (2)

where A is horizontal area and the integral is taken onan isobar. In Figure 3, we show results for the integralabove taken over the entire globe (top row) and over the

dayside (second row) and nightside (third row) alone.We use the time-resolved output from the last 150 daysof our simulation to calculate the amplitude of variability,defined as

∆Trms

Trms,max=Trms,max − Trms,min

Trms,max. (3)

In Equation (3), Trms,max and Trms,min are the maxi-mum and minimum global root-mean-square tempera-tures (i.e., the maximum and minimum of Trms) over thelast 150 days of the simulation, respectively.

It is possible that significant small-scale variability isnot captured by the global-scale variability metrics dis-cussed above. As a result, we calculate an additionalmetric for the root-mean-square temperature variabilityof local regions as a function of time over the last 150days of the simulation:

Trms,loc(λ, φ, p) =

√∫ [T (λ, φ, p, t)− T (λ, φ, p)

]2dt

tint.

(4)In Equation (4), Trms,loc is the local root-mean-squaretemperature variability (which is a function of longitudeλ, latitude φ, and pressure p), T is the spatial and time-dependent temperature, T is the time-average of the tem-perature over the last 150 days of simulation time, andtint = 150 days is the time interval over which we takethe integral. Then, we calculate the spatial root-mean-square of Trms,loc as in Equation (2) to obtain a globallyaveraged metric for the time-variability in localized re-gions, and normalize by horizontal root-mean-square ofthe time-averaged temperature T (which we term Tav).We show this local variability metric calculated from oursuite of simulations in the bottom row of Figure 3.

In general, the time-variability in global-rms tempera-ture (top row of Figure 3) is of the order 0.1− 1%. Thelocal variability (bottom row of Figure 3) is generallylarger than the global variability, and there is always sig-nificant local variability even in cases with strong drag.However, in a global-mean, much of this local variabilitycancels out, producing the relatively small global tem-perature variability. Additionally, the global variabilityis slightly less than the local variability in equatorial re-gions shown in Figure 1, as there is much less variabilityat higher latitudes.

The variability amplitude generally increases with de-creasing pressure. This is likely because the day-to-nightforcing that drives the equatorial jet increases with de-creasing pressure due to the shorter radiative timescalesat lower pressures (Komacek & Showman 2016). Thevariability amplitude on the dayside and nightside sepa-rately (second and third rows of Figure 3, respectively)are generally greater than in the global-average, point-ing toward cancellation between the two in the globalmean. For example, at any given time, hotter-than-average daysides will be accommpanied by cooler-than-average nightsides. Additionally, the variability is gen-erally larger on the nightside than on the dayside, as asimilar absolute amplitude of temperature variability isrelatively larger on the colder nightside.

3.1.2. Wind speeds

Page 7: ATEX style emulateapj v. 12/16/11 - arXiv

Time Variability in Hot Jupiter Atmospheres 7

t = tfin t = tfin 150 days t = tfin 75 days(a) (b) (c)

Latitude (°)

Pres

sure

(Bar

s)

−80 −60 −40 −20 0 20 40 60 80

10−3

10−2

10−1

100

101

1020

1000

2000

3000

4000

Latitude (°)

Pres

sure

(Bar

s)

0

0

0 0 0 0

0

0

0

0

0

0

0

0 0

0

0

0

0

0

0

0

0

00

0

0

0

1

1 1

1

1

1 1

1

1

1 11

1

1

11

111

1

1

1

2

2

2

2

22 2

2 22

2 2

2

2

−80 −60 −40 −20 0 20 40 60 80

10−3

10−2

10−1

100

101

102

−3

−2

−1

0

1

2

3

Latitude (°)

Pres

sure

(Bar

s)

0

0

0

0

0

0 0

0

0

0

0

0

0

0

0

0 0

0

0

0

0

0

0

0

0

00

0 0

0

11

1

11

1

1

1

1

1

1

1

11

1

1

1

1

1

1

1

2

2

2

2

2

2 22 2

22

2

22

2

2

23

−80 −60 −40 −20 0 20 40 60 80

10−3

10−2

10−1

100

101

102

−3

−2

−1

0

1

2

3u [ms1]

<latexit sha1_base64="N4MX780LLWsBnpPI1JmzvKqKGzY=">AAACDXicbZDLSsNAFIYn9VbrLerSzWAV3FgSEXRZdOOygr1AEstkOm2HzkzCzEQoIT6AG1/FjQtF3Lp359s4aSNo6w8DH/85hznnD2NGlXacL6u0sLi0vFJeraytb2xu2ds7LRUlEpMmjlgkOyFShFFBmppqRjqxJIiHjLTD0WVeb98RqWgkbvQ4JgFHA0H7FCNtrK59AP0QyTTJ7j2fIz2UPOXZD6nsNj12s6BrV52aMxGcB7eAKijU6Nqffi/CCSdCY4aU8lwn1kGKpKaYkaziJ4rECI/QgHgGBeJEBenkmgweGqcH+5E0T2g4cX9PpIgrNeah6cz3VLO13Pyv5iW6fx6kVMSJJgJPP+onDOoI5tHAHpUEazY2gLCkZleIh0girE2AFROCO3vyPLROaq5Tc69Pq/WLIo4y2AP74Ai44AzUwRVogCbA4AE8gRfwaj1az9ab9T5tLVnFzC74I+vjG+lGnLU=</latexit><latexit sha1_base64="N4MX780LLWsBnpPI1JmzvKqKGzY=">AAACDXicbZDLSsNAFIYn9VbrLerSzWAV3FgSEXRZdOOygr1AEstkOm2HzkzCzEQoIT6AG1/FjQtF3Lp359s4aSNo6w8DH/85hznnD2NGlXacL6u0sLi0vFJeraytb2xu2ds7LRUlEpMmjlgkOyFShFFBmppqRjqxJIiHjLTD0WVeb98RqWgkbvQ4JgFHA0H7FCNtrK59AP0QyTTJ7j2fIz2UPOXZD6nsNj12s6BrV52aMxGcB7eAKijU6Nqffi/CCSdCY4aU8lwn1kGKpKaYkaziJ4rECI/QgHgGBeJEBenkmgweGqcH+5E0T2g4cX9PpIgrNeah6cz3VLO13Pyv5iW6fx6kVMSJJgJPP+onDOoI5tHAHpUEazY2gLCkZleIh0girE2AFROCO3vyPLROaq5Tc69Pq/WLIo4y2AP74Ai44AzUwRVogCbA4AE8gRfwaj1az9ab9T5tLVnFzC74I+vjG+lGnLU=</latexit><latexit sha1_base64="N4MX780LLWsBnpPI1JmzvKqKGzY=">AAACDXicbZDLSsNAFIYn9VbrLerSzWAV3FgSEXRZdOOygr1AEstkOm2HzkzCzEQoIT6AG1/FjQtF3Lp359s4aSNo6w8DH/85hznnD2NGlXacL6u0sLi0vFJeraytb2xu2ds7LRUlEpMmjlgkOyFShFFBmppqRjqxJIiHjLTD0WVeb98RqWgkbvQ4JgFHA0H7FCNtrK59AP0QyTTJ7j2fIz2UPOXZD6nsNj12s6BrV52aMxGcB7eAKijU6Nqffi/CCSdCY4aU8lwn1kGKpKaYkaziJ4rECI/QgHgGBeJEBenkmgweGqcH+5E0T2g4cX9PpIgrNeah6cz3VLO13Pyv5iW6fx6kVMSJJgJPP+onDOoI5tHAHpUEazY2gLCkZleIh0girE2AFROCO3vyPLROaq5Tc69Pq/WLIo4y2AP74Ai44AzUwRVogCbA4AE8gRfwaj1az9ab9T5tLVnFzC74I+vjG+lGnLU=</latexit><latexit sha1_base64="N4MX780LLWsBnpPI1JmzvKqKGzY=">AAACDXicbZDLSsNAFIYn9VbrLerSzWAV3FgSEXRZdOOygr1AEstkOm2HzkzCzEQoIT6AG1/FjQtF3Lp359s4aSNo6w8DH/85hznnD2NGlXacL6u0sLi0vFJeraytb2xu2ds7LRUlEpMmjlgkOyFShFFBmppqRjqxJIiHjLTD0WVeb98RqWgkbvQ4JgFHA0H7FCNtrK59AP0QyTTJ7j2fIz2UPOXZD6nsNj12s6BrV52aMxGcB7eAKijU6Nqffi/CCSdCY4aU8lwn1kGKpKaYkaziJ4rECI/QgHgGBeJEBenkmgweGqcH+5E0T2g4cX9PpIgrNeah6cz3VLO13Pyv5iW6fx6kVMSJJgJPP+onDOoI5tHAHpUEazY2gLCkZleIh0girE2AFROCO3vyPLROaq5Tc69Pq/WLIo4y2AP74Ai44AzUwRVogCbA4AE8gRfwaj1az9ab9T5tLVnFzC74I+vjG+lGnLU=</latexit>

log10

u [ms1]

<latexit sha1_base64="8EpQWFVC5r78NvH1k/9eut5rUSE=">AAACM3icbVDLSsNAFJ34tr6qLt0MFkEXlkQEXRZ1Ia4UbBWaWCbTm3ZwJgkzN0IJ8Zvc+CMuBHGhiFv/wWltwdeBgcM55zL3njCVwqDrPjlj4xOTU9Mzs6W5+YXFpfLySsMkmeZQ54lM9GXIDEgRQx0FSrhMNTAVSrgIrw/7/sUNaCOS+Bx7KQSKdWIRCc7QSq3yia8YdrXKZdIpWrnnFr6ECDf9I5DIqB8ynWcFvaXNUVAVI2aKq3zbKwJfi04Xt1rlilt1B6B/iTckFTLEaav84LcTnimIkUtmTNNzUwxyplFwCUXJzwykjF+zDjQtjZkCE+SDmwu6YZU2jRJtX4x0oH6fyJkypqdCm+yva357ffE/r5lhtB/kIk4zhJh/fRRlkmJC+wXSttDAUfYsYVwLuyvlXaYZR1tzyZbg/T75L2nsVD236p3tVmoHwzpmyBpZJ5vEI3ukRo7JKakTTu7II3khr8698+y8Oe9f0TFnOLNKfsD5+AQWTaxg</latexit><latexit sha1_base64="8EpQWFVC5r78NvH1k/9eut5rUSE=">AAACM3icbVDLSsNAFJ34tr6qLt0MFkEXlkQEXRZ1Ia4UbBWaWCbTm3ZwJgkzN0IJ8Zvc+CMuBHGhiFv/wWltwdeBgcM55zL3njCVwqDrPjlj4xOTU9Mzs6W5+YXFpfLySsMkmeZQ54lM9GXIDEgRQx0FSrhMNTAVSrgIrw/7/sUNaCOS+Bx7KQSKdWIRCc7QSq3yia8YdrXKZdIpWrnnFr6ECDf9I5DIqB8ynWcFvaXNUVAVI2aKq3zbKwJfi04Xt1rlilt1B6B/iTckFTLEaav84LcTnimIkUtmTNNzUwxyplFwCUXJzwykjF+zDjQtjZkCE+SDmwu6YZU2jRJtX4x0oH6fyJkypqdCm+yva357ffE/r5lhtB/kIk4zhJh/fRRlkmJC+wXSttDAUfYsYVwLuyvlXaYZR1tzyZbg/T75L2nsVD236p3tVmoHwzpmyBpZJ5vEI3ukRo7JKakTTu7II3khr8698+y8Oe9f0TFnOLNKfsD5+AQWTaxg</latexit><latexit sha1_base64="8EpQWFVC5r78NvH1k/9eut5rUSE=">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</latexit><latexit sha1_base64="8EpQWFVC5r78NvH1k/9eut5rUSE=">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</latexit>

Fig. 4.— Maps showing the typical zonal-mean zonal wind velocity and the characteristic time-variation in the zonal-mean zonal windfor the simulation with Teq = 1500 K and τdrag = ∞. (a) Zonal-mean zonal wind as a function of pressure and latitude at the endof the simulation, t = tfin = 2000 days. A superrotating jet, centered on the equator, is seen with a peak eastward wind speed of≈ 4750 m s−1. (b,c) Map of the logarithmic change in zonal-mean zonal wind speed between t = tfin − 150 days and t = tfin (b) andbetween t = tfin − 75 days and t = tfin (c). Contours are over-plotted in log10 units, with contour labels of 0, 1, 2, and 3 corresponding tochanges in wind speed of 1, 10, 100, and 1000 m s−1. In the area of the superrotating jet, the change in the zonal-mean zonal wind speedis largest, with characteristic amplitudes of ∼ 1− 10%.

Along with the variability in temperature discussedabove, our model hot Jupiter atmospheres are signifi-cantly time-variable in wind speeds. Figure 1 (bottom)shows the zonal wind speed and its variability over thelast 150 days of the simulation with Teq = 1500 K andτdrag = ∞. The zonal wind itself peaks at the equatorand decreases towards higher latitudes. The variabilityin wind speed is also largest near the equator, with muchof the variability occurring in the region where the equa-torial jet is present. The location of the variability inzonal wind speed is similar to that in temperature (recallthe top panel of Figure 1), and has a similar maximumamplitude of ∼ 2%.

Figure 4 shows the time-variability in the zonal-meanzonal wind speed (east-west average of the east-westwind), analogous to Figure 1 but analyzing how the windspeed variability depends on pressure. The greatest vari-ability in zonal-mean zonal wind speed occurs near thebase of the jet, at a pressure of several bars where the lo-cal variability can reach ∼ 10%. Note that regions closerto the expected photospheric level (shown in Figure 1)show smaller variability, of order 1% in the zonal-meanzonal wind. The regions on the flanks of the jet (northand south of the equatorial regions) show the least time-variability. These regions poleward of the jet are wherethe eddy-momentum flux that drives the equatorial jetoriginates (Showman & Polvani 2011, Showman et al.2015).

The variability in wind speed discussed above mani-fests itself as global-scale variability with a characteristicamplitude of 1−10%. In Figure 5, we show the amplitudeof the rms wind speed variability at different pressuresfrom our suite of simulations with varying equilibriumtemperature and drag strength. We calculate the rmswind speed at a given pressure as

Urms(p) =

√∫(u2 + v2)dA

A, (5)

where as before u and v are the zonal and meridionalcomponents of wind speed, respectively, and the inte-gral is taken over the globe. We calculate the variabilityamplitude of rms wind speed as in Equation (3), simply

substituting Urms for Trms.In Figure 5, one can see that the global-rms wind speed

variability is generally significantly larger than the tem-perature variability. As for the temperature variability,we see that the τdrag = 106 s case exhibits relatively largevariability compared to other drag timescales. Unlike forthe temperature variability, the variability in wind speedcan still be significant for τdrag = 105 s, and is larger thanthe variability for the case with τdrag =∞ s at pressures≥ 80 mbars and equilibrium temperatures ≥ 2000 K.This is because the τdrag = 105 s model lies betweenthe regimes with circulation dominated by an equatorialsuperrotating jet and with circulation flowing from day-to-night. This leads to a forced-damped oscillation inthe wind speeds, with the forcing due to the day-nightirradiation difference driving wave propagation and thedamping due to the applied drag.

In general, the cases with weaker drag (τdrag & 105 s)show much greater variability in global-rms wind speeds.In Section 4, we will discuss how high-resolution spectro-scopic observations of hot Jupiters may be able to observetime-variability in Doppler shifts of spectral lines.

3.2. Observable time-variability

To calculate how time-variability may be observable,we compute the outgoing radiated infrared flux at eachtime-step in each of our simulations. Figure 6 showsmaps of the outgoing flux and its change over time forour simulation with Teq = 1500 K and τdrag = ∞. Onecan see that the outgoing flux pattern looks similar to thetemperature pattern shown in Figure 1, with an east-ward shift of the flux maximum and a large day-nightflux contrast. Similar to the temperature variability,the flux variability is largest near the equator. How-ever, the flux variability can be significantly larger thanthe temperature variability, reaching up to ∼ 10% inlocalized regions. This is because variability in emit-ted flux is generally larger than that in temperature.Consider a blackbody, where F = σT 4, with σ theStefan-Boltzmann constant. The relative change in fluxdF/F = 4σT 3dT/F = 4σT 3dT/

(σT 4

)= 4dT/T . For a

blackbody, the flux variability is 4 times larger than thetemperature variability.

Page 8: ATEX style emulateapj v. 12/16/11 - arXiv

8 Komacek & Showman

1000 1500 2000 2500 3000Equilibrium Temperature (Kelvin)

0

5

10

15

∆U

rms/U

rms,

max

[%]

p = 1 barτdrag =∞τdrag = 107sec

τdrag = 106sec

τdrag = 105sec

τdrag = 104sec

τdrag = 103sec

1000 1500 2000 2500 3000Equilibrium Temperature (Kelvin)

0

2

4

6

8

10

p = 80 mbars

1000 1500 2000 2500 3000Equilibrium Temperature (Kelvin)

0.0

2.5

5.0

7.5

10.0

12.5

p = 1 mbar

Fig. 5.— Normalized amplitude of variability in the root-mean-square horizontal wind speed at given pressure levels over the last 150 daysof simulation time. The wind speed variability is shown for all for simulations with varying τdrag = 103 −∞ s and Teq = 1000 − 3000 Kat three different pressures: p = 1 bar, 80 mbars, and 1 mbar. Similar to the time-variability in the RMS of temperature (Figure 3),the variability in wind speed generally (but not always) increases with increasing equilibrium temperature and decreasing pressure level.Additionally, deep in the atmosphere (p ≥ 80 mbars), the variability is strongest for τdrag = 106 s.

t = tfin t = tfin 150 days t = tfin 75 days(a) (b) (c)F [Wm2]<latexit sha1_base64="arQDKaGWADaVo5JZPTQ/PbQXS1A=">AAACBnicbZDLSgMxFIYz9VbrbdSlCMEiuLHMFEGXRUFcVrAXmI4lk6ZtaJIZkoxQhnHjxldx40IRtz6DO9/GTDuCtv4Q+PjPOeScP4gYVdpxvqzCwuLS8kpxtbS2vrG5ZW/vNFUYS0waOGShbAdIEUYFaWiqGWlHkiAeMNIKRhdZvXVHpKKhuNHjiPgcDQTtU4y0sbr2/uW91+FIDyVPWukP8fQ2Oa6mftcuOxVnIjgPbg5lkKvetT87vRDHnAiNGVLKc51I+wmSmmJG0lInViRCeIQGxDMoECfKTyZnpPDQOD3YD6V5QsOJ+3siQVypMQ9MZ7anmq1l5n81L9b9Mz+hIoo1EXj6UT9mUIcwywT2qCRYs7EBhCU1u0I8RBJhbZIrmRDc2ZPnoVmtuE7FvT4p187zOIpgDxyAI+CCU1ADV6AOGgCDB/AEXsCr9Wg9W2/W+7S1YOUzu+CPrI9vEXOZfA==</latexit><latexit sha1_base64="arQDKaGWADaVo5JZPTQ/PbQXS1A=">AAACBnicbZDLSgMxFIYz9VbrbdSlCMEiuLHMFEGXRUFcVrAXmI4lk6ZtaJIZkoxQhnHjxldx40IRtz6DO9/GTDuCtv4Q+PjPOeScP4gYVdpxvqzCwuLS8kpxtbS2vrG5ZW/vNFUYS0waOGShbAdIEUYFaWiqGWlHkiAeMNIKRhdZvXVHpKKhuNHjiPgcDQTtU4y0sbr2/uW91+FIDyVPWukP8fQ2Oa6mftcuOxVnIjgPbg5lkKvetT87vRDHnAiNGVLKc51I+wmSmmJG0lInViRCeIQGxDMoECfKTyZnpPDQOD3YD6V5QsOJ+3siQVypMQ9MZ7anmq1l5n81L9b9Mz+hIoo1EXj6UT9mUIcwywT2qCRYs7EBhCU1u0I8RBJhbZIrmRDc2ZPnoVmtuE7FvT4p187zOIpgDxyAI+CCU1ADV6AOGgCDB/AEXsCr9Wg9W2/W+7S1YOUzu+CPrI9vEXOZfA==</latexit><latexit sha1_base64="arQDKaGWADaVo5JZPTQ/PbQXS1A=">AAACBnicbZDLSgMxFIYz9VbrbdSlCMEiuLHMFEGXRUFcVrAXmI4lk6ZtaJIZkoxQhnHjxldx40IRtz6DO9/GTDuCtv4Q+PjPOeScP4gYVdpxvqzCwuLS8kpxtbS2vrG5ZW/vNFUYS0waOGShbAdIEUYFaWiqGWlHkiAeMNIKRhdZvXVHpKKhuNHjiPgcDQTtU4y0sbr2/uW91+FIDyVPWukP8fQ2Oa6mftcuOxVnIjgPbg5lkKvetT87vRDHnAiNGVLKc51I+wmSmmJG0lInViRCeIQGxDMoECfKTyZnpPDQOD3YD6V5QsOJ+3siQVypMQ9MZ7anmq1l5n81L9b9Mz+hIoo1EXj6UT9mUIcwywT2qCRYs7EBhCU1u0I8RBJhbZIrmRDc2ZPnoVmtuE7FvT4p187zOIpgDxyAI+CCU1ADV6AOGgCDB/AEXsCr9Wg9W2/W+7S1YOUzu+CPrI9vEXOZfA==</latexit><latexit sha1_base64="arQDKaGWADaVo5JZPTQ/PbQXS1A=">AAACBnicbZDLSgMxFIYz9VbrbdSlCMEiuLHMFEGXRUFcVrAXmI4lk6ZtaJIZkoxQhnHjxldx40IRtz6DO9/GTDuCtv4Q+PjPOeScP4gYVdpxvqzCwuLS8kpxtbS2vrG5ZW/vNFUYS0waOGShbAdIEUYFaWiqGWlHkiAeMNIKRhdZvXVHpKKhuNHjiPgcDQTtU4y0sbr2/uW91+FIDyVPWukP8fQ2Oa6mftcuOxVnIjgPbg5lkKvetT87vRDHnAiNGVLKc51I+wmSmmJG0lInViRCeIQGxDMoECfKTyZnpPDQOD3YD6V5QsOJ+3siQVypMQ9MZ7anmq1l5n81L9b9Mz+hIoo1EXj6UT9mUIcwywT2qCRYs7EBhCU1u0I8RBJhbZIrmRDc2ZPnoVmtuE7FvT4p187zOIpgDxyAI+CCU1ADV6AOGgCDB/AEXsCr9Wg9W2/W+7S1YOUzu+CPrI9vEXOZfA==</latexit>

F [Wm2]<latexit sha1_base64="yekBIWssPcITHXfFzWxZrV1Ysso=">AAACDXicbZDLSsNAFIYnXmu9RV26GayCG0tSBF0WFXFZwV4giWUynbRDZ5IwMxFKiA/gxldx40IRt+7d+TZO2gja+sPAx3/OYc75/ZhRqSzry5ibX1hcWi6tlFfX1jc2za3tlowSgUkTRywSHR9JwmhImooqRjqxIIj7jLT94Xleb98RIWkU3qhRTDyO+iENKEZKW11z370gTCF4ee+4HKmB4Gk7+yGe3aZHtczrmhWrao0FZ8EuoAIKNbrmp9uLcMJJqDBDUjq2FSsvRUJRzEhWdhNJYoSHqE8cjSHiRHrp+JoMHminB4NI6BcqOHZ/T6SISznivu7M95TTtdz8r+YkKjj1UhrGiSIhnnwUJAyqCObRwB4VBCs20oCwoHpXiAdIIKx0gGUdgj198iy0alXbqtrXx5X6WRFHCeyCPXAIbHAC6uAKNEATYPAAnsALeDUejWfjzXiftM4ZxcwO+CPj4xsGSJwo</latexit><latexit sha1_base64="yekBIWssPcITHXfFzWxZrV1Ysso=">AAACDXicbZDLSsNAFIYnXmu9RV26GayCG0tSBF0WFXFZwV4giWUynbRDZ5IwMxFKiA/gxldx40IRt+7d+TZO2gja+sPAx3/OYc75/ZhRqSzry5ibX1hcWi6tlFfX1jc2za3tlowSgUkTRywSHR9JwmhImooqRjqxIIj7jLT94Xleb98RIWkU3qhRTDyO+iENKEZKW11z370gTCF4ee+4HKmB4Gk7+yGe3aZHtczrmhWrao0FZ8EuoAIKNbrmp9uLcMJJqDBDUjq2FSsvRUJRzEhWdhNJYoSHqE8cjSHiRHrp+JoMHminB4NI6BcqOHZ/T6SISznivu7M95TTtdz8r+YkKjj1UhrGiSIhnnwUJAyqCObRwB4VBCs20oCwoHpXiAdIIKx0gGUdgj198iy0alXbqtrXx5X6WRFHCeyCPXAIbHAC6uAKNEATYPAAnsALeDUejWfjzXiftM4ZxcwO+CPj4xsGSJwo</latexit><latexit sha1_base64="yekBIWssPcITHXfFzWxZrV1Ysso=">AAACDXicbZDLSsNAFIYnXmu9RV26GayCG0tSBF0WFXFZwV4giWUynbRDZ5IwMxFKiA/gxldx40IRt+7d+TZO2gja+sPAx3/OYc75/ZhRqSzry5ibX1hcWi6tlFfX1jc2za3tlowSgUkTRywSHR9JwmhImooqRjqxIIj7jLT94Xleb98RIWkU3qhRTDyO+iENKEZKW11z370gTCF4ee+4HKmB4Gk7+yGe3aZHtczrmhWrao0FZ8EuoAIKNbrmp9uLcMJJqDBDUjq2FSsvRUJRzEhWdhNJYoSHqE8cjSHiRHrp+JoMHminB4NI6BcqOHZ/T6SISznivu7M95TTtdz8r+YkKjj1UhrGiSIhnnwUJAyqCObRwB4VBCs20oCwoHpXiAdIIKx0gGUdgj198iy0alXbqtrXx5X6WRFHCeyCPXAIbHAC6uAKNEATYPAAnsALeDUejWfjzXiftM4ZxcwO+CPj4xsGSJwo</latexit><latexit sha1_base64="yekBIWssPcITHXfFzWxZrV1Ysso=">AAACDXicbZDLSsNAFIYnXmu9RV26GayCG0tSBF0WFXFZwV4giWUynbRDZ5IwMxFKiA/gxldx40IRt+7d+TZO2gja+sPAx3/OYc75/ZhRqSzry5ibX1hcWi6tlFfX1jc2za3tlowSgUkTRywSHR9JwmhImooqRjqxIIj7jLT94Xleb98RIWkU3qhRTDyO+iENKEZKW11z370gTCF4ee+4HKmB4Gk7+yGe3aZHtczrmhWrao0FZ8EuoAIKNbrmp9uLcMJJqDBDUjq2FSsvRUJRzEhWdhNJYoSHqE8cjSHiRHrp+JoMHminB4NI6BcqOHZ/T6SISznivu7M95TTtdz8r+YkKjj1UhrGiSIhnnwUJAyqCObRwB4VBCs20oCwoHpXiAdIIKx0gGUdgj198iy0alXbqtrXx5X6WRFHCeyCPXAIbHAC6uAKNEATYPAAnsALeDUejWfjzXiftM4ZxcwO+CPj4xsGSJwo</latexit>

Longitude (°)

Latit

ude

(° )

−150 −100 −50 0 50 100 150

−80

−60

−40

−20

0

20

40

60

80

1

2

3

4

5

6

x 105

Longitude (°)

Latit

ude

(° )

−150 −100 −50 0 50 100 150

−80

−60

−40

−20

0

20

40

60

80

−6

−4

−2

0

2

4

6x 104

Longitude (°)

Latit

ude

(° )

−150 −100 −50 0 50 100 150

−80

−60

−40

−20

0

20

40

60

80

−6

−4

−2

0

2

4

6x 104

Fig. 6.— Maps of flux at the end of the simulation and the characteristic time-variation in the emitted flux for the simulation withTeq = 1500 K and τdrag = ∞. (a) Emitted flux at the top of the atmosphere at the end of the simulation, t = tfin = 2000 days. Theflux map closely follows the temperature map shown in Figure 1. (b,c) Map of the change in emitted flux between t = tfin − 150 days andt = tfin (b) and between t = tfin − 75 days and t = tfin (c). The largest flux variations occur at low latitudes, similar to the variability intemperature and wind speed. However, the amplitude of flux variability in localized regions can be up to ∼ 10%, significantly larger thanthe local temperature variability.

Given that we find potentially large-amplitude fluxvariability, we next determine to what extent this fluxvariability might be observable by constructing modelphase curves. To do so, we calculate the flux that theEarth-facing hemisphere of the model planet radiates to-ward Earth at each orbital phase using the method de-scribed in Section 3.3 of Komacek et al. (2017). Thisobservable flux is an area-weighted average of the fluxover a hemisphere that migrates in longitude over time.In this way we construct a full-phase light curve, or phasecurve, for each orbital period of the last 150 days in eachsimulation.

In Figure 7, we show the amplitude of variability inthe light curve from each simulation. In general, theflux variability at any given orbital phase is small, at the. 1 − 3% level over a broad range in equilibrium tem-perature and drag strength. As we found for the time-variability of atmospheric temperature and wind speed,the variability in emitted flux increases with increasingdrag timescale and increasing equilibrium temperatureup to τdrag = 106 s, then decreases at even larger values ofτdrag. Additionally, as before the case with τdrag = 106 s

has the largest amplitude of flux variability.Using these phase curves, we then calculate three key

parameters: (1) the flux emerging from the hemispherecentered on the substellar point, which is the flux theplanet radiates at secondary eclipse, (2) the phase curveamplitude, measured as the difference between the max-imum and minimum flux of the phase curve, and (3) thephase curve offset, which is the offset in orbital phasebetween the phase at which the Earth-facing hemisphereof the planet is brightest and secondary eclipse. We showthe resulting secondary eclipse variation versus time for asubset of simulations with Teq = 1500, 2000, and 2500 Kand τdrag = 106 s and τdrag = ∞ in Figure 8. Then,we calculate the normalized amplitude of variability ofthese three parameters for all simulations, using the samemethod as in Equation (3), and plot the amplitude oftime-variability in the secondary eclipse flux, phase curveamplitude, and phase offset in Figure 9.

3.2.1. Secondary eclipse depth

Figure 8 shows the time-evolution of our simulated sec-ondary eclipse depths for a subset of cases with Teq =

Page 9: ATEX style emulateapj v. 12/16/11 - arXiv

Time Variability in Hot Jupiter Atmospheres 9

0.0

0.2

0.4

0.6

0.8

1.0

1.2

Flu

x[W

m2

]

107 Teq = 3000 K

drag =1drag = 107sec

drag = 106sec

drag = 105sec

drag = 104sec

drag = 103sec

0 60 120 180 240 300 360Orbital phase ()

103

104

105

Flu

xV

aria

bili

ty[W

m2

]

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

Flu

x[W

m

2]

105 Teq = 1000 K

0 60 120 180 240 300 360Orbital phase ()

102

103

Flu

xV

aria

bili

ty[W

m

2]

1

2

3

4

5

6

7

Flu

x[W

m

2]

105 Teq = 1500 K

0 60 120 180 240 300 360Orbital phase ()

102

103

104F

lux

Var

iabili

ty[W

m

2] 0.0

0.5

1.0

1.5

2.0

Flu

x[W

m

2]

106 Teq = 2000 K

0 60 120 180 240 300 360Orbital phase ()

102

103

104

Flu

xV

aria

bili

ty[W

m

2]

0

1

2

3

4

5

Flu

x[W

m

2]

106 Teq = 2500 K

0 60 120 180 240 300 360Orbital phase ()

103

104

105

Flu

xV

aria

bili

ty[W

m

2] 0.0

0.2

0.4

0.6

0.8

1.0

1.2

Flu

x[W

m

2]

107 Teq = 3000 K

0 60 120 180 240 300 360Orbital phase ()

103

104

105

Flu

xV

aria

bili

ty[W

m

2]

Fig. 7.— Variability in the infrared phase curves produced by our suite of double-grey general circulation models. The top panels showthe simulated emitted flux from the planet as a function of orbital phase, with the width of each line representing the maximum variation inemitted flux at each orbital phase over the last 150 days of simulation time. The bottom panels show the maximum variability in emittedflux (the width of the lines in the top panels) as a function of orbital phase. Secondary eclipse (shown by the dashed vertical line) occursat an orbital phase of 180, and transit occurs at an orbital phase of 0. Here the flux is a hemispheric average of the flux emitted towardthe line of sight of the observer. As seen for the variability in intrinsic physical parameters (see Figures 3 and 5), the light curve variabilityis largest for the case with τdrag = 106 s and smallest in the cases with strong drag, τdrag . 104 s. Additionally, the variability is largestnear secondary eclipse, due to significant variability in the phase curve offset (see Figure 9).

1500, 2000, and 2500 K and τdrag =∞ and τdrag = 106 s.Generally, we find that the secondary eclipse depths formodels with τdrag = ∞ are variable at the . 0.5%level, and the secondary eclipse depths for models withτdrag = 106 s are variable at the ∼ 1% level. For thecase with τdrag = ∞, the secondary eclipse depth vari-ability is markedly larger for higher Teq, but there is lessof a difference in secondary eclipse depth with Teq for thecases with τdrag = 106 s. This secondary eclipse depthvariability is only quasi-periodic, with small-amplitudevariability occurring on a timescale of days and larger-amplitude variability occurring on timescales of weeksto months. The Fourier transform of each light curveshows that there are not individual dominant variabilityfrequencies. This differs from the results of Showman

et al. (2009), who found that the simulated variability insecondary eclipse depth was regular in time. However,the Fourier transform shows that the variability is low-frequency, with significant power at frequencies below0.5 cycles per day. Note that the Fourier transform ofthe variability in phase curve amplitude and offset (notshown) shows similar power at low frequencies.

Though it may not be detectable in the near future,we also show the variability of the emitted flux from thenightside of the planet during transit in Figure 8. Thevariability in the emitted flux from the nightside of theplanet is significant, and is larger than the secondaryeclipse variability for both τdrag = ∞ and τdrag = 106 sby a factor of ∼ 2. Similar to the secondary eclipse depthvariability, there is no single dominant frequency of the

Page 10: ATEX style emulateapj v. 12/16/11 - arXiv

10 Komacek & Showman

1900 1910 1920 1930 1940 1950Time [Earth days]

0.9925

0.9950

0.9975

1.0000

1.0025

1.0050

1.0075

Rel

ativ

enig

htsi

de

emit

ted

flux

drag =1Transit

0.0 0.2 0.4 0.6 0.8 1.0Frequency [cycles per day]

0.05

0.00

0.05F

FT

1900 1910 1920 1930 1940 1950Time [Earth days]

0.996

0.998

1.000

1.002

1.004

Rel

ativ

ese

condar

yec

lipse

dep

th drag =1Secondary eclipse

0.0 0.2 0.4 0.6 0.8 1.0Frequency [cycles per day]

0.05

0.00

0.05

FF

T

1900 1910 1920 1930 1940 1950

0.996

0.998

1.000

1.002

1.004

Rel

ativ

ese

condar

yec

lipse

dep

th drag =1Secondary eclipse

Teq = 1500 K

Teq = 2000 K

Teq = 2500 K

1900 1910 1920 1930 1940 19500.9925

0.9950

0.9975

1.0000

1.0025

1.0050

1.0075

Rel

ativ

enig

htsi

de

emit

ted

flux

drag =1Transit

1900 1910 1920 1930 1940 1950Time [Earth days]

0.98

0.99

1.00

1.01

Rel

ativ

ese

condar

yec

lipse

dep

th drag = 106 s

1900 1910 1920 1930 1940 1950Time [Earth days]

0.97

0.98

0.99

1.00

1.01

1.02

1.03

Rel

ativ

enig

htsi

de

emit

ted

flux drag = 106 s

1900 1910 1920 1930 1940 1950Time [Earth days]

0.98

0.99

1.00

1.01

Rel

ativ

ese

condar

yec

lipse

dep

th

drag = 106 s

0.0 0.2 0.4 0.6 0.8 1.0Frequency [cycles per day]

0.1

0.0

0.1

FF

T

1900 1910 1920 1930 1940 1950Time [Earth days]

0.97

0.98

0.99

1.00

1.01

1.02

1.03

Rel

ativ

enig

htsi

de

emit

ted

flux

drag = 106 s

0.0 0.2 0.4 0.6 0.8 1.0Frequency [cycles per day]

0.5

0.0

FF

T

Fig. 8.— Normalized secondary eclipse depth vs. time (left) and in-transit emitted flux vs. time for simulations with τdrag = ∞ (top)

and τdrag = 106 s (bottom). Curves show results from simulations with Teq = 1500, 2000, and 2500 K, with darker lines corresponding tocooler temperatures. Simulated data are plotted twice for every Earth day, with 50 Earth days of time shown in each plot. The fast Fouriertransform of each light curve is shown below the plot. Note that here we simply output the flux variations from hemispheres centered uponthe substellar point (left) and anti-stellar point (right) rather than computing more detailed mock secondary eclipse and transit spectra. Ingeneral, for the cases with τdrag =∞ the secondary eclipse depth and in-transit emitted flux amplitudes are . 0.5%, though they increase

with increasing equilibrium temperature. For the case with τdrag = 106 s, both amplitudes are markedly larger, at the ∼ 1% level forsecondary eclipse depth and few percent level for in-transit emitted flux. The Fourier transforms of the light curves show no clear peak,but in general the variability is low-frequency, with significant power at frequencies . 0.4 cycles per day.

Page 11: ATEX style emulateapj v. 12/16/11 - arXiv

Time Variability in Hot Jupiter Atmospheres 11

1000 1500 2000 2500 3000

10−1

100

Sec

ond

ary

eclip

sed

epth

vari

abili

ty[%

]

(a)

τdrag =∞τdrag = 107 s

τdrag = 106 s

τdrag = 105 s

τdrag = 104 s

τdrag = 103 s

1000 1500 2000 2500 3000

10−2

10−1

100

Ph

ase

amp

litu

de

vari

abili

ty[%

] (b)

1000 1500 2000 2500 3000Equilibrium Temperature (Kelvin)

0

2

4

6

Ph

ase

offse

tva

riab

ility

[]

(c)

Fig. 9.— Amplitude of variability in key light curve parame-ters over the last 150 days of simulation time. (a) Percent vari-ation in the flux from the hemisphere centered on the substel-lar point for simulations with varying τdrag = 103 − ∞ s andTeq = 1000 − 3000 K. The amplitude of emergent substellar fluxvariation generally increases with increasing τdrag, but is largest

for τdrag = 106 s. Generally, the amplitude of variation is . 2%.(b) Percent variation in the phase curve amplitude for the sameset of simulations. The variation is 1-2 orders of magnitude largerfor simulations with τdrag ≥ 105 s. (c) Absolute variation in phasecurve offset, in degrees longitude, for the same set of simulations.Simulations with strong drag (τdrag ≤ 104 s) show no variation in

phase curve offset, and the τdrag = 107 s and τdrag =∞ case showthe same variability amplitude in phase curve offset.

variability in emitted flux from the nightside hemisphere.Figure 9(a) shows the amplitude of the variability in

secondary eclipse depth for all simulations. We find thatthe secondary eclipse depths of our model hot Jupitersare variable at the . 2% level. The variability in sec-ondary eclipse depth generally increases with increas-ing equilibrium temperature, from ∼ 0.3% − 1% overTeq = 1000 − 3000 K for the case with no applied drag.Similar to the temperature and wind speed variability,the case with τdrag = 106 s shows significant variabilityin secondary eclipse depth. It is notable that the caseswith strong drag (τdrag . 105 s) show significantly lessvariability than cases with weaker drag. Additionally,there is not a strong dependence of the time-variabilityin secondary eclipse depth with equilibrium temperaturefor cases with τdrag . 104 s. We will return to our resultsfor secondary eclipse time-variability in Section 4 to dis-cuss how secondary eclipse variability may be observablewith JWST.

3.2.2. Phase curve amplitude

Figure 9(b) shows the variability in phase curve am-plitude for our grid of simulations. The variability inphase curve amplitude is comparable to the variabilityin secondary eclipse depth, as the variability in phasecurve amplitude is dominated by the variability in thehottest hemisphere. In general, for simulations withτdrag & 105 s we find that the phase curve amplitude vari-ability is ∼ 0.3%−3%. However, for strong τdrag . 104 sthe phase curve amplitude is much less variable, at the. 0.1% level for Teq = 1000 K and as small as 0.01%for Teq & 2000 K. Additionally, unlike the secondaryeclipse depth, the phase curve amplitude variability isstill significant for the case with τdrag = 105 s. This maybe related to how the case with τdrag = 105 s lies in anintermediate regime between an atmosphere dominatedby a superrotating jet and an atmosphere dominated bydayside-to-nightside flow (see Figure 5 of Komacek et al.2017). As a result, the superrotating equatorial jet isslowed, but not as much as it is with stronger drag. Thismay lead to an oscillation due to the similarity betweenthe wave driving and damping timescales (which are both∼ 105 s). This residual variability in the wave pattern isseen through the phase curve amplitude variability.

3.2.3. Phase curve offset

Figure 9(c) shows the variability in phase curve offsetfor our suite of simulations. For cases with weak drag(τdrag & 105 s), the variability in phase offset is signif-icant, up to ≈ 5. For the cases with very weak drag(τdrag & 107 s), the phase curve offset variability reaches≈ 3 at Teq = 3000 K. As we found for all other observ-able properties, the τdrag = 106 s case shows the largestvariability in phase curve offset, up to ≈ 5. We do notfind significant variability for the cases with strong drag(τdrag . 104 s). This is both because these cases withstrong drag do not have a significant phase offset to be-gin with due to their weak wind speeds and because thestrong drag then damps any potential variability in thephase curve offset.

4. DISCUSSION

4.1. Mechanism of the variability

Page 12: ATEX style emulateapj v. 12/16/11 - arXiv

12 Komacek & Showman

drag = 1 drag = 106 s

T

T(

)[K

]<latexit sha1_base64="KcvyXRUzLXQXqlx1TWKg2gjcjA0=">AAACCXicbVDLSsNAFJ3UV62vqEs3o0WoC0sigroruhHcVGhsIQllMpm0QycPZiZCCXHrxl9x40LFrX/gzr9x0mahrQcGDufcw9x7vIRRIQ3jW6ssLC4tr1RXa2vrG5tb+vbOnYhTjomFYxbznocEYTQilqSSkV7CCQo9Rrre6Krwu/eECxpHHTlOiBuiQUQDipFUUl/f78Bj6HiIZ5284TAV9NHRg+2ESA55mN3kbl+vG01jAjhPzJLUQYl2X/9y/BinIYkkZkgI2zQS6WaIS4oZyWtOKkiC8AgNiK1ohEIi3GxySg4PleLDIObqRRJO1N+JDIVCjENPTRYrilmvEP/z7FQG525GoySVJMLTj4KUQRnDohfoU06wZGNFEOZU7QrxEHGEpWqvpkowZ0+eJ9ZJ86Jp3J7WW5dlG1WwBw5AA5jgDLTANWgDC2DwCJ7BK3jTnrQX7V37mI5WtDKzC/5A+/wBoBiZ0g==</latexit><latexit sha1_base64="KcvyXRUzLXQXqlx1TWKg2gjcjA0=">AAACCXicbVDLSsNAFJ3UV62vqEs3o0WoC0sigroruhHcVGhsIQllMpm0QycPZiZCCXHrxl9x40LFrX/gzr9x0mahrQcGDufcw9x7vIRRIQ3jW6ssLC4tr1RXa2vrG5tb+vbOnYhTjomFYxbznocEYTQilqSSkV7CCQo9Rrre6Krwu/eECxpHHTlOiBuiQUQDipFUUl/f78Bj6HiIZ5284TAV9NHRg+2ESA55mN3kbl+vG01jAjhPzJLUQYl2X/9y/BinIYkkZkgI2zQS6WaIS4oZyWtOKkiC8AgNiK1ohEIi3GxySg4PleLDIObqRRJO1N+JDIVCjENPTRYrilmvEP/z7FQG525GoySVJMLTj4KUQRnDohfoU06wZGNFEOZU7QrxEHGEpWqvpkowZ0+eJ9ZJ86Jp3J7WW5dlG1WwBw5AA5jgDLTANWgDC2DwCJ7BK3jTnrQX7V37mI5WtDKzC/5A+/wBoBiZ0g==</latexit><latexit sha1_base64="KcvyXRUzLXQXqlx1TWKg2gjcjA0=">AAACCXicbVDLSsNAFJ3UV62vqEs3o0WoC0sigroruhHcVGhsIQllMpm0QycPZiZCCXHrxl9x40LFrX/gzr9x0mahrQcGDufcw9x7vIRRIQ3jW6ssLC4tr1RXa2vrG5tb+vbOnYhTjomFYxbznocEYTQilqSSkV7CCQo9Rrre6Krwu/eECxpHHTlOiBuiQUQDipFUUl/f78Bj6HiIZ5284TAV9NHRg+2ESA55mN3kbl+vG01jAjhPzJLUQYl2X/9y/BinIYkkZkgI2zQS6WaIS4oZyWtOKkiC8AgNiK1ohEIi3GxySg4PleLDIObqRRJO1N+JDIVCjENPTRYrilmvEP/z7FQG525GoySVJMLTj4KUQRnDohfoU06wZGNFEOZU7QrxEHGEpWqvpkowZ0+eJ9ZJ86Jp3J7WW5dlG1WwBw5AA5jgDLTANWgDC2DwCJ7BK3jTnrQX7V37mI5WtDKzC/5A+/wBoBiZ0g==</latexit>

Longitude (°)

Tim

e (d

ays)

−100 0 1001970

1975

1980

1985

1990

1995

−10

−5

0

5

10

15

Longitude (°)

Tim

e (d

ays)

−100 0 1001970

1975

1980

1985

1990

1995

−100

−50

0

50

100

150

Fig. 10.— Hovmoller diagrams of temperature deviations from the time-averaged temperature, averaged over latitudes of ±15, as afunction of longitude (λ) at a given time at a pressure of 80 mbar for simulations with Teq = 1500 K and τdrag = 106 s and ∞. Thediagrams show only 25 days of simulation time, from t = 1970− 1995 days. Multiple wave trains generated on the western hemisphere ofthe dayside and propagating eastward are visible in the τdrag =∞ panel. Additionally, westward propagation of waves is visible near the

eastern terminator. The characteristic variability for the case with τdrag = 106 s is an order of magnitude larger than for the case withτdrag = ∞, and this variability is much noisier. As a result, it is difficult to visually discern wave propagation for this case, as expectedfrom the time-series analysis of secondary eclipse depth shown in Figure 8.

Though in this work we do not perform a detailed anal-ysis of the simulated time-variability, in this section wesummarize some of the key features of the variability andspeculate on possible mechanisms that cause this vari-ability. Recall from the temperature variability maps inFigure 1 that the bulk of the time-variability is confinedto the equatorial regions, within a latitudinal band be-tween ±15. Additionally, the variability is largest nearboth the western and eastern terminators. Notably, thevariability at the western terminator coincides with theposition of a large temperature contrast between the coolnightside and hot dayside.

To analyze the time-variability in greater detail, wereturn to our analysis of how the equatorial tempera-ture pattern changes as a function of time. Figure 10shows Hovmoller diagrams for the equatorial tempera-ture deviation from the time-averaged temperature as afunction of longitude for simulations with Teq = 1500 Kand τdrag = 106 s and ∞. This is similar to Figure 2,but here the temperature deviation is relative to the lo-cal time-average of temperature instead of the longitude-average. Additionally, we have focused on only 25 daysof the simulation, in order to visually discern features ofthe time-variability. One can see that the amplitude ofvariability for the case with τdrag = 106 s is generally anorder of magnitude larger than the case with τdrag =∞.Additionally, the spatial distribution of such variabilityis somewhat different. The strongest variability for thecase with τdrag = 106 s occurs near the western termi-nator, with additional variability eastward of the easternterminator where there is vertical advection due to theGill pattern. However, the strongest variability for thecase with τdrag =∞ occurs just eastward of the westernterminator. This variability occurs over a much smallerrange of longitudes than in the case with τdrag = 106 s,likely due to the much lower-amplitude variability in the

case with τdrag = ∞. In general, the variability forthe case with τdrag = 106 s is fairly noisy, with indi-vidual wave trains difficult to see. However, the casewith τdrag =∞ does have clear wave propagation, as wediscuss next.

Now we speculate on the wave generation mechanismusing the results shown in Figure 10 for the case withτdrag = ∞. The temperature pattern appears to havetilts in longitude-time phase space such that the tilts areeastward (with increasing time) eastward of the westernterminator and westward (again, with increasing time)westward of the western terminator. Overall, the vari-ability is relatively noisy, with periodicity difficult to dis-cern. However, waves can be seen over a broad range oflongitudes from −50 to 180. Both eastward and west-ward propagating modes can be prominently seen, su-perimposed together. The eastward-propagating wavespropagate from their generation region on the westernhalf of the dayside to the substellar point in a timespan ofa few days, corresponding to relatively low phase speedsof ∼ 600 m s−1. The westward propagating waves thatoccur at eastward longitudes have similar phase speeds.The basic characteristics of the eastward-propagatingwaves are reminiscent of Kelvin waves (Holton & Hakim2013): they propagate eastward, are confined to equato-rial regions, and have relatively low phase speeds. Addi-tionally, one might expect that the westward propagatingbranch of waves may be equatorial Rossby waves, whichare also planetary-scale waves with relatively low phasespeeds that propagate westward.

To examine this in more detail, we compare the ∼600 m s−1 propagation speed of both the eastwardand westward propagating waves to that of Rossby andKelvin waves. This propagation speed is broadly con-

Page 13: ATEX style emulateapj v. 12/16/11 - arXiv

Time Variability in Hot Jupiter Atmospheres 13

sistent with the expected phase speed of Rossby waves2,but is a factor of two smaller than that of Kelvin waves3

when ignoring the Doppler shift due to the backgroundjet. Including the Doppler shift of the background jet(with a zonal-mean equatorial speed of ∼ 4000 m s−1),one would expect the phase speeds of both the Doppler-shifted Kelvin and Rossby waves to be an order of mag-nitude larger than that found in Figure 10. As a result,it is unclear whether Rossby and/or Kelvin waves arethe cause of the equatorial variability in our simulations.An alternative is that the equatorial waves are com-posed of mixed Rossby-gravity modes, which is consistentwith the anti-symmetry of the temperature perturbationsacross the equator found in Figure 1. Most likely, theequatorial time-variability in our simulations is causedby a combination of mixed Rossby-gravity, Kelvin, andRossby waves, which cannot be discerned without moredetailed wavenumber-frequency analysis.

Note that the ∼ 10 day generation frequency and∼ 20 day longitudinal propagation timescale (across thecircumference of the planet) of the waves are not dis-similar to the ∼ 40 day periodicity of variability seenby Showman et al. (2009), given the stark differences inour modeling approaches. With higher-resolution simu-lations, it may be possible to do a more detailed analysisof the wavenumber-frequency spectrum of hot Jupiterflows, akin to that in the Earth tropical dynamic litera-ture (Wheeler & Kiladis 1999). Such an analysis has beenperformed on simulations of brown dwarfs and Jupiter-like planets (Showman et al. 2019, Young et al. 2019).A similar analysis is outside the scope of this work, butwould be beneficial for determining the wave types caus-ing time-variability in hot Jupiter atmospheres.

Lastly, we speculate on the process generating theseplanetary-scale waves. It appears that the waves origi-nate near the western terminator, at the location wherethe temperature increases sharply going from nightsideto dayside. This potential wave generation region cor-responds to a hydraulic jump or acoustic shock on thewestern limb that has been noted previously by manyauthors (e.g., Showman et al. 2009, Heng 2012, Pernaet al. 2012, Fromang et al. 2016). Note that though ournumerical simulations do not explicitly resolve acousticshocks, the hydraulic jumps seen in this work and that ofShowman et al. (2009) are analogous features. If they arepresent, shocks may then act to cause stirring that gen-erates planetary-scale propagating waves. For example,small fluctuations in the longitude of the shock over timewould act as an impulse that excites Kelvin waves whichwould then propagate eastward away from the shock fea-ture. Future work building upon that of Fromang et al.

2 The phase speed of a Rossby wave in the limit of zero back-ground wind is approximately c ≈ −β/k2, where β is the latitudi-nal gradient in the Coriolis parameter and k is horizontal wavenum-ber. Using a planetary radius and rotation rate appropriate for HD209458b, a planetary wavenumber of 16 (estimated from the casewith τdrag = ∞), and evaluating the phase speed at the equator,

we find c ≈ −600 m s−1.3 In the limit of long vertical wavelength, the phase speed of a

Kelvin wave without background flow is c ≈ 2NH, where N is theBrunt-Vaisala frequency and H is the scale height (Showman et al.

2013). In the vertically isothermal limit, c ≈ R√T/cp, where

R is the specific gas constant, T is temperature, and cp is thespecific heat. For parameters appropriate for the simulation withTeq = 1500 K, c ≈ 1250 m s−1.

(2016) using simulations that allow for shocks and theresulting wave generation can shed further light on thecause of time-variability in hot Jupiter atmospheres.

4.2. Doppler wind speeds: effects of varying equilibriumtemperature and drag timescale

High-resolution spectroscopy has constrained theterminator-averaged blueshift for HD 209458b (Snellenet al. 2010) and both the terminator-averaged andthe terminator-separated (i.e., separated into those forthe eastern and western limbs) Doppler shifts for HD189733b (Louden & Wheatley 2015, Wyttenbach et al.2015, Brogi et al. 2016). These observations probe rel-atively high altitudes in the atmosphere, correspondingto pressures of ∼ 1 mbar or less. Under the assump-tion that HD 189733b is tidally locked to its host star,Louden & Wheatley (2015) separated the Doppler shiftson the leading and trailing limbs of the planet during atransit event. Louden & Wheatley found that the windspeeds were redshifted at the leading (western) limb andblueshifted at the trailing (eastern) limb, exactly what isexpected from the presence of a superrotating jet (Show-man & Guillot 2002, Showman & Polvani 2011, Kemp-ton & Rauscher 2012, Showman et al. 2013, Zhang et al.2017, Flowers et al. 2019).

In Figure 11 (top), we show the simulated line-of-sightwind speeds at a pressure of 1 mbar at both the east-ern and western terminators and their average for eachof our simulations varying equilibrium temperature anddrag strength. Specifically, we show the time-mean ofthe latitudinally-averaged zonal wind speeds on the ter-minators. We define positive wind speeds as away fromthe observer (i.e., westward on the eastern limb and east-ward on the western limb) and negative wind speeds astoward the observer (i.e., eastward on the eastern limband westward on the western limb).

We find that the wind speeds on the eastern limb arealways negative (toward the observer) and increase inabsolute magnitude with increasing equilibrium temper-ature and decreasing drag timescale, as expected fromthe simulated net Doppler wind speeds of previous stud-ies (Kempton & Rauscher 2012, Showman et al. 2013,Kempton et al. 2014). The wind speeds on the westernlimb are negative for all simulations with τdrag ≤ 107 sand generally increase in magnitude with increasing equi-librium temperature. On the western limb, models withno drag experience positive winds (i.e., away from theobserver leading to a Doppler redshift) for modest Teqof 1000-1500 K. This is the regime where the equatorialjet is most coherent with longitude (see Figure 1d, andFigure 5 of Komacek et al. 2017) and thus this is the sig-nature of the equatorial jet blowing away from observer.Note that the wind speeds on the western limb do notchange monotonically with drag strength, as the fastestwinds at the western limb are for the simulations with in-termediate drag strengths. The average of the two limbsis always negative, with the fastest winds toward the ob-server for simulations with τdrag = 106 s.

We note that our simulations with τdrag = ∞ areconsistent with the observations of Louden & Wheat-ley (2015) in that at the low equilibrium temperatures of1000-1500 K relevant for HD 189733b, the eastern limbshows a net blueshift, the western limb a net redshift,and their combination shows a net blueshift (see their

Page 14: ATEX style emulateapj v. 12/16/11 - arXiv

14 Komacek & Showman

1000 1500 2000 2500 3000

−6

−4

−2

0

Lin

e−

of−

sigh

tw

ind

spee

d[k

ms−

1] Eastern (trailing) limb

1000 1500 2000 2500 3000

−6

−4

−2

0

Western (leading) limb

τdrag =∞τdrag = 107 s

τdrag = 106 s

τdrag = 105 s

τdrag = 104 s

τdrag = 103 s

1000 1500 2000 2500 3000

−6

−4

−2

0

Average of east and west limbs

1000 1500 2000 2500 3000Equilibrium Temperature (Kelvin)

10−5

10−4

10−3

10−2

10−1

100

Win

dsp

eed

vari

abili

ty[k

ms−

1]

1000 1500 2000 2500 3000Equilibrium Temperature (Kelvin)

10−5

10−4

10−3

10−2

10−1

100

1000 1500 2000 2500 3000Equilibrium Temperature (Kelvin)

10−5

10−4

10−3

10−2

10−1

100

Fig. 11.— Simulated line-of-sight wind speed at a pressure of 1 mbar (top) and the maximum time-variability in these wind speeds overthe last 150 days of the simulation (bottom). Specifically, the top three panels show the time-mean over the last 150 days of the averagewind speed toward the observer across the eastern and western terminator and their average. The bottom three panels show the maximumvariability in these average wind speeds (i.e., the difference between the maximum and minimum line-of-sight wind speeds) over the last150 days of simulation time. The top panels share a linear scale, with positive wind speeds corresponding to redshifts (motion away fromthe observer), while negative wind speeds correspond to blueshifts (motion toward the observer). The bottom panels share a logarithmicscale. Wind speeds at the eastern limb are always negative, and increase in absolute magnitude with increasing equilibrium temperatureand decreasing drag strength. For simulations with high equilibrium temperature and strong drag, wind speeds at the western limb arenormally negative, but do not monotonically increase in absolute magnitude with decreasing drag strength. The average of the wind speedon the east and west limbs is always negative and increases in absolute magnitude with increasing equilibrium temperature. The variabilityin terminator wind speeds is large for cases with weak drag (τdrag ≥ 106 s) and is normally larger on the western limb than on the eastern

limb. The variability in terminator-averaged wind speeds can be a few kilometers per second for cases with weak drag (τdrag ≥ 106 s),potentially observable with current high-resolution spectroscopic techniques.

Figure 3). Our models predict that hotter hot Jupitersmay have a western limb that (like the eastern limb) be-comes blueshifted.

4.3. Future prospects for observing time-variability

4.3.1. Light curves

As shown in Section 3.2, we expect that the secondaryeclipse depths of hot Jupiters are time-variable at the. 2% level. Notably, we show in Figure 9 that for at-mospheres with weak drag (τdrag & 107 s), the secondaryeclipse variability increases with increasing equilibriumtemperature, from ∼ 0.3% at Teq = 1000 K to ∼ 1%at Teq = 3000 K. This variability would not be de-tectable with Spitzer given current upper limits (Agolet al. 2010, Kilpatrick et al. 2019). As a result, we nextdetermine if the observable variability from our simulatedhot Jupiters would be detectable with JWST.

For JWST, we assume a pessimistic noise floor of≈ 30 ppm with NIRCam (Greene et al. 2016) and con-sider a planet-to-star flux ratio of ≈ 0.002 in this wave-

length range (relevant for an HD 209458b-like planetwith Teq = 1500 K). With the above assumptions, weestimate that JWST will be able to detect planetary fluxvariability on the order of & 1.5%. Assuming that theplanet to star flux ratio increases as a blackbody, weexpect that for hot Jupiters with Teq & 2000 K time-variability in secondary eclipse depth in the near-infraredon the order of & 0.1% will likely be detectable. Our sim-ulations show that variability amplitudes that could beobservable with JWST are possible for planets with weakτdrag & 106 s.

We expect potentially significant variability in thephase curve offset, up to 3 for the hottest hot Jupiterswithout atmospheric drag and potentially up to ≈ 5

for hot Jupiters with moderate τdrag = 105 − 106 s.Though the variability we model is generally not as largeas the variability that Armstrong et al. (2016) and Jack-son et al. (2019) found for HAT-P-7b and Kepler-76bin visible wavelengths and Bell et al. (2019) found in theinfrared for WASP-12b, the variability in phase curve off-

Page 15: ATEX style emulateapj v. 12/16/11 - arXiv

Time Variability in Hot Jupiter Atmospheres 15

1850 1875 1900 1925 1950 1975Time [days]

−4

−3

−2

−1

0

Lin

e-of

-sig

htw

ind

spee

d[k

ms−

1]

Teq = 2000 K, τdrag =∞Western limb

Eastern limb

Average

Fig. 12.— Time-dependence of the line-of-sight wind speeds at1 mbar from the simulation with Teq = 2000 K and τdrag = ∞.The Doppler shift at the western limb can reverse sign from ablueshift to a redshift over short timescales. The amplitude ofvariability is much smaller on the eastern limb, which is alwaysblueshifted. As a result, the terminator-averaged wind speeds arealways blueshifted. Three other GCM experiments that we per-formed show similar behavior, with the Doppler shift at the west-ern limb reversing sign. High-resolution spectroscopic techniqueswhich measure the Doppler shift at the ingress and egress sepa-rately may find a change in sign of the wind speed at the leading(western) limb due to purely hydrodynamic effects.

set that we find is still signifiant and potentially observ-able with multiple phase curves. Additionally, note thatour simulations do not include magnetic effects, whichmay greatly affect the phase curve offset at temperatures& 1500 K, and we do not include clouds (see Section 4.4for further discussion). Note that though full-phase lightcurves are a very time-intensive measurement (Crossfield2015, Parmentier & Crossfield 2017), it is possible thatfuture dedicated missions (e.g., ARIEL) may be able toobserve multiple phase curves of the same object in orderto constrain atmospheric time-variability.

4.3.2. Doppler wind speeds

It is possible that the Doppler wind speeds presentedin Section 4.2 may vary in time, providing informationabout the variability of the underlying atmospheric cir-culation. In Figure 11 (bottom), we show the ampli-tude of time-variability in the line-of-sight wind speedsat the eastern and western terminators, along with theirsum. We find that the variability is orders-of-magnitudelarger for simulations with τdrag ≥ 106 s than for thosewith τdrag ≤ 104 s. Generally, for simulations withτdrag ≥ 106 s the wind speed variability on the easternlimb is smaller than that on the western limb.

Figure 12 shows that for the case with Teq = 2000 Kand no drag, the line-of-sight wind speed averaged overthe western limb at a pressure of 1 mbar can reversefrom blueshifted to redshifted over short timescales. Thisreversal of the line-of-sight wind speed at the westernlimb occurs in 4 separate experiments, all with weakdrag. The specific experiments that show this behav-ior have equilibrium temperatures and drag timescalesof Teq = 1000 K and τdrag = 107 s, Teq = 1500 Kand τdrag = 107 s, Teq = 2000 K and τdrag = ∞, andTeq = 2500 K and τdrag = ∞. This behavior occursbecause the wind speed at the western limb is on aver-age near zero, due to a cancellation between the eastward

(blueshifted) superrotating jet and westward (redshifted)day-to-night flow at higher latitudes. As a result, similar-amplitude variability in the wind speed as found in othersimulations can lead to a reversal in the line-of-sight winddirection.

For all simulations with Teq ≥ 1500 K and τdrag ≥107 s, the wind-speed variability on the western limband for the average of the eastern and western limb canbe & 1 km s. This variability is both significant rel-ative to the wind speeds themselves and may be ob-servable, given that the current state-of-the-art uncer-tainty on limb-separated wind speeds is ≈ ±1.5 km s−1,with the uncertainty for the limb-averaged wind speeds≈ ±0.5 km s−1 (Louden & Wheatley 2015). As a re-sult, we expect that future observations of multiple tran-sit events with high-resolution spectroscopic techniquesmay enable the detection of time-variability of Dopplershifts in high-resolution spectra.

4.4. Other mechanisms to induce atmosphericvariability

In this work, we undertook a first exploration ofhow time-variability in hot Jupiter atmospheres varieswith planetary parameters using a simplified three-dimensional GCM. Notably, we did not include non-greyradiative transfer nor the effects of clouds, the latter ofwhich is known to greatly affect phase curves and emer-gent spectra of hot Jupiters (Demory et al. 2013, Arm-strong et al. 2016, Heng 2016, Lee et al. 2016, Parmentieret al. 2016, Stevenson 2016, Lines et al. 2018). Usinga GCM with passive tracers, Parmentier et al. (2013)showed that the tracer abundance can vary significantlymore than the temperature field itself. As a result, thiscould allow for significant variability in the cloud dis-tribution in a GCM that includes clouds. Additionally,it is possible that small variations in temperature couldlead to large variations in the saturation vapor pressureof condensible species via the Clausius-Clapeyron rela-tion (Pierrehumbert 2010, Ch. 2.6). This implies thatsmall temperature variations could cause large variationsin condensible vapor abundances, which could in turncause significant cloud variability. Such cloud variabilitywould then alter the three-dimensional structure of ra-diative heating and cooling, which would then feed backinto the temperature structure and perhaps allow for en-hanced variability in the temperature field itself relativeto what our idealized models in this work show. As a re-sult, there is significant research yet to be done into howclouds affect time-variability in hot Jupiter atmospheres,including both their intrinsic time variability and the ob-servational consequences of this variability in spectra andphase curves.

As discussed in Section 1, magnetic effects can have asignificant effect on emergent properties of hot Jupiters.Most importantly, magnetic effects can cause the equato-rial jet to reverse from eastward to westward, oscillatingin time (Rogers & Komacek 2014, Rogers 2017). Thiswould cause resulting oscillations in the sign of the phasecurve offset and could cause the Doppler shifts in transitingress and egress to reverse sign (Louden & Wheatley2015). Note that we found that in certain parameterregimes purely hydrodynamic variability can also causea reversal of the sign of the Doppler shift at the west-ern limb (see Figure 12), so further work is needed to

Page 16: ATEX style emulateapj v. 12/16/11 - arXiv

16 Komacek & Showman

predict the impact of magnetic effects on high-resolutiontransmission spectra. In this work, we did not considertime-variability due to magnetic effects, though it is ex-pected that for hot Jupiters with Teq & 1400 K magneticeffects will be important (Menou 2012, Rogers & Show-man 2014, Rogers & Komacek 2014).

Additionally, our simulations do not capture non-hydrostatic instabilities associated with vertical shear ofthe horizontal winds, which have been shown to causetime-variability in the speed of the equatorial jet (Fro-mang et al. 2016). As a result, our predictions for thevariability in wind speeds are lower limits, and it is pos-sible that globally averaged wind speed variability atlevels of > 10% is achievable. In general, we did notinclude any parameterization for the drag timescale atnear-photospheric levels changing as a function of timeand space, as it could in an atmosphere with magneticeffects and/or vertical shear instabilities. We hope thatthis paper motivates future work exploring how more re-alistic drag parameterizations induce atmospheric time-variability in hot Jupiters.

5. CONCLUSIONS

We have shown that hot Jupiter atmospheres are time-variable at potentially observable levels, even when ig-noring the potential variability due to magnetic effects,vertical shear instabilities, and clouds. Below, we sum-marize the key ways in which this time-variability maybe manifest in observations, along with our predictionsfor how terminator wind speeds vary with planetary pa-rameters.

1. Hot Jupiter atmospheres are time-variable at the∼ 0.1−1% level in global-average, dayside-average,and nightside-average temperature and at the ∼1 − 10% level in global-average wind speeds. Thisvariability depends strongly on pressure, withlarger-amplitude variability generally occurring atlower pressures. With repeated transit observa-tions using high-resolution spectrographs, the &1 km s−1 variability in Doppler wind speeds atthe terminator predicted from our simulations withτdrag ≥ 106 s may be observationally detectable.

2. We calculate model Doppler wind speeds at a pres-sure of 1 mbar for the western (leading) and eastern(trailing) limbs in our suite of GCMs. We find thatthe eastern limb always shows a Doppler blueshift

due to winds. Meanwhile, the western limb shows asmall redshift in simulations without drag and withequilibrium temperatures ≤ 1500 K, but shows ablueshift in most simulations with stronger dragand/or larger values of incident stellar flux. Ad-ditionally, the time-variability in Doppler windspeeds can lead to redshift-to-blueshift oscillationson the western limb. As a result, even hot Jupiterswith a superrotating equatorial jet may not showa Doppler wind signature with an eastern limbblueshift and western limb redshift if they havelarge equilibrium temperatures and/or have signif-icant atmospheric drag.

3. Without considering magnetic effects and verticalshear instabilities, we expect that the secondaryeclipse depths of hot Jupiters are variable at the. 2% level. For atmospheres without strong ap-plied drag (with τdrag & 107 s), the secondaryeclipse depth variability increases with increasingequilibrium temperature, from ∼ 0.3%− 1% as Teqincreases from 1000 − 3000 K. For atmosphereswith strong drag (with τdrag . 105 s), the sec-ondary eclipse depth variability is always . 0.3%.Variability in the secondary eclipse depth of hotplanets with weak atmospheric drag is potentiallyobservable with JWST.

4. We expect that hot Jupiter phase curves are alsosignificantly time-variable. Most notably, we ex-pect that the phase offset, which has already beenshown to be variable from both optical and infraredphase curves, should be variable in the infraredif the characteristic drag timescale τdrag & 105 s.This variability in phase offset increases with in-creasing equilibrium temperature, and can be ≈3 − 5 for the hottest planets with weak atmo-spheric drag. We also predict that the phase curveamplitude is variable at the ∼ 0.1%− 1% level foratmospheres with τdrag & 105 s, with this variabil-ity in phase curve amplitude largely independentof equilibrium temperature.

T.D.K. acknowledges funding from the 51 Pegasi bFellowship in Planetary Astronomy sponsored by theHeising-Simons Foundation.

REFERENCES

Adcroft, A., Hill, C., Campin, J., Marshall, J., & Heimbach, P.2004, Monthly Weather Review, 132, 2845

Agol, E., Cowan, N., Knutson, H., Deming, D., Steffen, J., Henry,G., & Charbonneau, D. 2010, The Astrophysical Journal, 721,1861

Armstrong, D., de Mooij, E., Barstow, J., Osborn, H. P., Blake,J., & Fereshteh Saniee, N. 2016, Nature Astronomy, 1, 4

Batygin, K. & Stanley, S. 2014, The Astrophysical Journal, 794,10

Bell, T., Zhang, M., Cubillos, P., Dang, L., Fossati, L., Todorov,K., Cowan, N., Deming, D., Dobbs-Dixon, I., Fortney, J.,Knutson, H., & Line, M. 2019, Monthly Notices of the RoyalAstronomical Society

Brogi, M., de Kok, R., Albrecht, S., Snellen, I., Birkby, J., &Schwarz, H. 2016, The Astrophysical Journal, 817, 106

Cho, J., Polichtchouk, I., & Thrastarson, H. 2015, MonthlyNotices of the Royal Astronomical Society

Crossfield, I. 2015, Publications of the astronomical society of thepacific, 127, 941

Demory, B., Gillon, M., Madhusudhan, N., & Queloz, D. 2016,Monthly Notices of the Royal Astronomical Society, 455, 2018

Demory, B., Wit, J. D., Lewis, N., Fortney, J., Zsom, A., Seager,S., Knutson, H., Heng, K., Madhusudhan, N., Gillon, M.,Barclay, T., Desert, J., Parmentier, V., & Cowan, N. 2013, TheAstrophysical Journal Letters, 776, L25

Flowers, E., Brogi, M., Rauscher, E., Kempton, E., & Chiavassa,A. 2019, The Astronomical Journal, 157, 209

Fromang, S., Leconte, J., & Heng, K. 2016, Astronomy andAstrophysics, 591, A144

Gill, A. 1980, Quarterly Journal of the Royal MeteorologicalSociety, 106, 447

Page 17: ATEX style emulateapj v. 12/16/11 - arXiv

Time Variability in Hot Jupiter Atmospheres 17

Greene, T., Line, M., Montero, C., Fortney, J., Lustig-Yaeger, J.,& Luther, K. 2016, The Astrophysical Journal, 817, 17

Hammond, M. & Pierrehumbert, R. 2018, The AstrophysicalJournal, 869, 65

Heng, K. 2012, The Astrophysical Journal Letters, 761, L1—. 2016, The Astrophysical Journal Letters, 826, L16Heng, K., Frierson, D., & Phillips, P. 2011, Monthly Notices of

the Royal Astronomical Society, 418, 2669Heng, K. & Showman, A. 2015, Annual Reviews in Earth and

Planetary Sciences, 43, 509Hindle, A., Bushby, P., & Rogers, T. 2019, The Astrophysical

Journal Letters, 872, L27Holton, J. & Hakim, G. 2013, An introduction to dynamic

meteorology, 5th edn. (Academic Press)Jackson, B., Adams, E., Sandidge, W., Kreyche, S., & Briggs, J.

2019, The Astronomical Journal, 157, 239Kaspi, Y. & Flierl, G. 2006, Journal of the Atmospheric Sciences,

64, 3177Kempton, E., Perna, R., & Heng, K. 2014, The Astrophysical

Journal, 795, 24Kempton, E. & Rauscher, E. 2012, The Astrophysical Journal,

751, 117Kilpatrick, B., Kataria, T., Lewis, N., Zellem, R., Henry, G.,

Cowan, N., de Wit, J., Fortney, J., Knutson, H., Seager, S.,Showman, A., & Tucker, G. 2019, arXiv e-prints:1904.02294

Koll, D. & Komacek, T. 2018, The Astrophysical Journal, 853,133

Komacek, T. & Showman, A. 2016, The Astrophysical Journal,821, 16

Komacek, T., Showman, A., & Parmentier, V. 2019, TheAstrophysical Journal, 881, 152

Komacek, T., Showman, A., & Tan, X. 2017, The AstrophysicalJournal, 835, 198

Kylling, A., Stamnes, K., & Tsay, S. 1995, Journal ofAtmospheric Chemistry, 21, 115

Lee, G., Dobbs-Dixon, I., Helling, C., Bognar, K., & Woitke, P.2016, Astronomy and Astrophysics, 594, A48

Lian, Y. & Showman, A. 2008, Icarus, 194, 597Lines, S., Mayne, N., Boutle, I., Manners, J., Lee, G., Helling, C.,

Drummond, B., Amundsen, D., Goyal, J., Acreman, D.,Tremblin, P., & Kerslake, M. 2018, Astronomy & Astrophysics,615, A97

Liu, B. & Showman, A. 2013, The Astrophysical Journal, 770, 42Liu, J., Goldreich, P., & Stevenson, D. 2008, Icarus, 196, 653Louden, T. & Wheatley, P. 2015, The Astrophysical Journal

Letters, 814, L24Matsuno, T. 1966, Journal of the Meteorological Society of

Japan, 44, 25Menou, K. 2012, The Astrophysical Journal, 745, 138Menou, K. & Rauscher, E. 2009, The Astrophysical Journal, 700,

887Parmentier, V. & Crossfield, I. The Exoplanet HandbookParmentier, V., Fortney, J., Showman, A., Morley, C., & Marley,

M. 2016, The Astrophysical Journal, 828, 22Parmentier, V., Showman, A., & Lian, Y. 2013, Astronomy and

Astrophysics, 558, A91Perez-Becker, D. & Showman, A. 2013, The Astrophysical

Journal, 776, 134Perna, R., Heng, K., & Pont, F. 2012, The Astrophysical Journal,

751, 59

Pierrehumbert, R. 2010, Principles of Planetary Climate(Cambridge: Cambridge University Press)

Polichtchouk, I. & Cho, J. Y.-K. 2012, Monthly Notices of theRoyal Astronomical Society, 424, 1307

Powell, D., Zhang, X., Gao, P., & Parmentier, V. 2018, TheAstrophysical Journal, 860, 18

Rauscher, E. & Menou, K. 2010, The Astrophysical Journal, 714,1334

—. 2012, The Astrophysical Journal, 750, 96Rogers, T. 2017, Nature Astronomy, 1Rogers, T. & Komacek, T. 2014, The Astrophysical Journal, 794,

132Rogers, T. & Mcelwaine, J. 2017, The Astrophysical Journal

Letters, 841, L26Rogers, T. & Showman, A. 2014, The Astrophysical Journal

Letters, 782, L4

Sanchez-Lavega, A. 2010, An introduction to planetaryatmospheres, 1st edn. (Boca Raton, FL: CRC Press)

Shapiro, R. 1971, Journal of the Atmospheric Sciences, 28, 523Showman, A., Cho, J., & Menou, K. Exoplanets, ed. S. Seager

(Tucson, AZ: University of Arizona Press)Showman, A., Cooper, C., Fortney, J., & Marley, M. 2008, The

Astrophysical Journal, 682, 559Showman, A., Fortney, J., Lewis, N., & Shabram, M. 2013, The

Astrophysical Journal, 762, 24Showman, A., Fortney, J., Lian, Y., Marley, M., Freedman, R.,

Knutson, H., & Charbonneau, D. 2009, The AstrophysicalJournal, 699, 564

Showman, A. & Guillot, T. 2002, Astronomy and Astrophysics,385, 166

Showman, A., Lewis, N., & Fortney, J. 2015, The AstrophysicalJournal, 801, 95

Showman, A. & Polvani, L. 2010, Geophysical Research Letters,37, L18811

—. 2011, The Astrophysical Journal, 738, 71Showman, A., Tan, X., & Zhang, X. 2019, The Astrophysical

Journal, 883, 4Snellen, I., de Kok, R., de Mooij, E., & Albrecht, S. 2010, Nature,

465, 1049Stamnes, K., Tsay, S., Wiscombe, W., & Jayaweera, K. 1988,

Applied Optics, 27, 2502Stevenson, K. 2016, The Astrophysical Journal Letters, 817, L16Tamburo, P., Mandell, A., Deming, D., & Garhart, E. 2018, The

Astronomical Journal, 155, 221Tan, X. & Komacek, T. 2019, arXiv e-prints:1910.01622Thorngren, D., Gao, P., & Fortney, J. 2019, The Astrophysical

Journal Letters, 884, L6Tsai, S., Dobbs-Dixon, I., & Gu, P. 2014, The Astrophysical

Journal, 793, 141Wheeler, M. & Kiladis, G. 1999, Journal of the Atmospheric

Sciences, 56, 374Williams, G. 1979, Journal of the Atmospheric Sciences, 36, 932Wyttenbach, A., Ehrenreich, D., Lovis, C., Udry, S., & Pepe, F.

2015, Astronomy & Astrophysics, 577, A62Young, R., Read, P., & Wang, Y. 2019, Icarus, 326, 225Zhang, J., Kempton, E., & Rauscher, E. 2017, The Astrophysical

Journal, 851, 84