A publishing partnership

The following article is Open access

The Optical Aurorae of Europa, Ganymede, and Callisto

, , , and

Published 2023 February 16 © 2023. The Author(s). Published by the American Astronomical Society.
, , Citation Katherine de Kleer et al 2023 Planet. Sci. J. 4 37 DOI 10.3847/PSJ/acb53c

Download Article PDF
DownloadArticle ePub

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

2632-3338/4/2/37

Abstract

The tenuous atmospheres of the Galilean satellites are sourced from their surfaces and produced by a combination of plasma-surface interactions and thermal processes. Even though they are thin, these atmospheres can be studied via their auroral emissions, and most work to date has focused on their aurorae at UV wavelengths. Here we present the first detections of the optical aurorae of Ganymede and Callisto, as well as detections of new optical auroral lines at Europa, based on observations of the targets over 10 Jupiter eclipses from 1998 to 2021 with Keck/HIRES. We present measurements of O i emission at 6300/6364, 5577, 7774, and 8446 Å and place upper limits on hydrogen at 6563 Å. These constitute the first detections of emissions at 7774 and 8446 Å at a planetary body other than Earth. The simultaneous measurement of multiple emission lines provides robust constraints on atmospheric composition. We find that the eclipse atmospheres of Europa and Ganymede are composed predominantly of O2, with average column densities of (4.1 ± 0.1) × 1014 cm−2 and (4.7 ± 0.1) × 1014 cm−2, respectively. We find weak evidence for H2O in Europa's bulk atmosphere at an H2O/O2 ratio of ∼0.25, and place only an upper limit on H2O in Ganymede's bulk atmosphere, corresponding to H2O/O2 < 0.6. The column density of O2 derived for Callisto is (4.0 ± 0.9) × 1015 cm−2 for an assumed electron density of 0.15 cm−3, but electron properties at Callisto's orbit are very poorly constrained.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

The tenuous atmospheres of the Galilean satellites Europa, Ganymede, and Callisto are composed of a combination of O2, O, H2O, and CO2 and are ultimately sourced from their surfaces. The detailed compositions of these atmospheres, along with their spatial distributions and temporal variabilities, thus provide information on their surface compositions and on surface modification processes such as sputtering.

The atmosphere of Europa was first detected via the 1356 and 1304 Å multiplets of atomic oxygen in emission using the Goddard High Resolution Spectrograph (GHRS) on the Hubble Space Telescope (HST; Hall et al. 1995); additional measurements of Europa's aurora and a first detection of the same auroral lines at Ganymede were subsequently made (Hall et al. 1998). The emission was attributed to electron-impact dissociative excitation of predominantly O2 for both Europa and Ganymede, with perhaps some contribution from atomic O especially in the case of Ganymede, on the basis of the emission ratio of the two multiplets. The strength of the emission lines corresponds to an O2 atmosphere of 2.4–14 × 1014 cm−2 at Europa and 1–10 × 1014cm−2 at Ganymede, using Voyager- and Galileo-based estimates for the density and temperature and the exciting electrons (Hall et al. 1995, 1998).

Various mechanisms have been proposed for the production of these tenuous atmospheres. Motivated by an early claim of a 1 μbar atmosphere at Ganymede (Carlson et al. 1973), Yung & McElroy (1977) proposed a sublimation-driven water atmosphere that produces a molecular O2 atmosphere due to photolysis of H2O, followed by the escape of the lighter hydrogen. However, this model predicts a surface density well above the subsequent Voyager 1 findings (Broadfoot et al. 1979; Wolff & Mendis 1983). Lanzerotti et al. (1978) proposed that Ganymede's atmosphere is produced by Jovian plasma bombardment, which sputters neutral molecules off the icy surface. The authors argued that this is a more efficient mechanism for producing O2, especially at the low end of plausible surface temperatures. Brown et al. (1980) then measured ion sputtering yields as a function of ice temperature, and more recent modeling work has calculated sputtering yields of O2 and H2 for Europa (Cassidy et al. 2013) and for icy satellites in general (Teolis et al. 2017). Thermal H2 and O2 products do not condense as efficiently as sputtered water molecules, and H2 readily escapes, leaving O2 as the primary constituent. Johnson et al. (1981, 1982) used these experiments to predict that a ∼ 1014 cm−2 atmosphere of O2 would be sustained at Ganymede and ∼1015 cm−2 at Europa, and also calculated an expected ∼2–100 × 1012 cm−2 of H2O across the three icy satellites. However, models subsequently showed that O2 yields from ion sputtering are insufficient to match the observed column densities, unless additional resputtering by freshly ionized O2 augments the flux from Jupiter's magnetosphere (Ip 1996). Saur et al. (1998) modeled the atmospheric sources and sinks, including plasma deflection, and found that a combination of suprathermal torus ions and thermal ions sputtering O2 from the surface could produce a stable atmosphere around the measured column densities. They also demonstrated that the resputtering mechanism could not contribute a significant amount.

Images of Europa and Ganymede in the UV O i multiplets (1304 and 1356 Å) with HST/STIS have yielded evidence for spatial variations across both satellites (Feldman et al. 2000; McGrath et al. 2013; Roth et al. 2016). For Europa, the spatial distribution of emissions exhibits some systematic trends, but is still not fully understood. A dusk-dawn asymmetry is observed in both the UV and optical auroral data (Roth et al. 2016; de Kleer & Brown 2019), which is consistent with simulations (Oza et al. 2019) and suggests a thermal role in the production of the O2 atmosphere (Oza et al. 2018; Johnson et al. 2019). For Ganymede, the aurora appears to behave analogously to Earth's auroral ovals, whereby electrons are accelerated into the near-surface region along field lines at the open-closed field line boundary (Feldman et al. 2000; Eviatar et al. 2001; McGrath et al. 2013). The oscillation amplitude of the ovals has been used as evidence to support the existence of a subsurface ocean on Ganymede (Saur et al. 2015). The brightness of Ganymede's auroral spots cannot be matched by models assuming the electron energies and densities at Ganymede's orbit, requiring either higher electron energies or higher-density lower-temperature electrons (Eviatar et al. 2001); the higher electron energies are attributed to local acceleration at Ganymede (Eviatar et al. 2001). The uncertainties on the electron population exciting the emissions permit O2 column densities in the range of 1–30 × 1014 cm−2 for Ganymede. In contrast, Europa's auroral emissions are consistent with excitation by thermal magnetospheric electrons and do not require local acceleration (Saur et al. 1998).

Recently, Roth (2021) and Roth et al. (2021) used auroral data sets to independently constrain the atomic O abundance in the atmospheres of Europa and Ganymede by measuring the resonant scattering component of the 1304 Å emission as the satellites passed through Jupiter's shadow. These data sets placed a tight upper limit on the O abundance, which then requires a new mechanism to explain the low 1356/1304 Å ratio (sometimes referred to as rγ (O i); e.g., Roth 2021) on the trailing hemispheres of these satellites. The proposed mechanism is a consistently present H2O atmosphere centered on the trailing hemisphere for Europa and on both hemispheres (but six times denser on the trailing) for Ganymede, which can be produced by sublimation in the case of Ganymede (as is also predicted on Callisto; Carberry Mogan et al. 2021) and by sputtering combined with sublimation of the fresh deposits of sputtered H2O in the case of Europa (Roth 2021; Teolis et al. 2017). The derived mixing ratios of H2O/O2 over the trailing hemisphere are in the 10–30 range for both satellites. The UV lines are significantly more sensitive to O2 than to H2O, and observations of H in addition to O, and/or observations of lines with higher intrinsic emission rates following electron impact on H2O, would strengthen the constraints on the presence and abundance of H2O.

To date, studies of the auroral emissions of Europa and Ganymede have been conducted almost exclusively in the UV. However, measurements of additional lines at different wavelengths can provide more robust constraints on the atmospheric composition and have the potential to reveal additional processes at work or even new atmospheric constituents. The optical auroral lines have been studied at Io for decades (e.g., Belton et al. 1996; Schmidt et al. 2023). On the icy Galilean satellites, whose atmospheres and orbital environments are both less populated than those of Io, the first published detection of the optical aurora was the 6300/6364 Å oxygen emission observed at Europa from HST/STIS and the Keck High Resolution Spectrograph (HIRES; de Kleer & Brown 2018, 2019). The data supported an O2 atmospheric composition and displayed a high level of temporal variability and spatial patchiness in addition to an overall asymmetry in the auroral brightness with more emission on the trailing/dusk side, consistent with the UV morphology observed by Roth et al. (2016).

Here we present observations of Europa, Ganymede, and Callisto in Jupiter eclipse over 10 eclipses in total between 1998 and 2021, taken with Keck/HIRES. The data cover wavelengths from 5000 to 9000 Å. We present measurements of the 6300/6364, 5577, 7774, and 8446 Å O i emissions from these satellites. All emission lines that are listed are detected on at least one occasion, and we present upper limits on dates when a given line is not detected, as well as on Hα 6563 Å on all dates. The observations and data reduction procedures are described in Section 2. The model used to interpret the data is expanded from that of de Kleer & Brown (2018) and is described in Section 3, including the Markov chain Monte Carlo (MCMC) retrieval algorithm. The derived auroral brightnesses are presented and discussed in Section 4, and conclusions are summarized in Section 5.

2. Observations and Data Reduction

Observations of Europa, Ganymede, and Callisto in Jupiter eclipse were obtained on four, four, and two occasions, respectively, between 1998 and 2021 using the HIRES instrument (Vogt et al. 1994) on the Keck I telescope at the summit of Maunakea in Hawaii. Observations in eclipse remove the reflected sunlight component, which would otherwise overwhelm the signal from the visible-wavelength aurora. One eclipse observation of Ganymede is from 1998, which used HIRES pre-upgrade when it had a single CCD and less extensive spectral coverage. Table 1 lists the observing parameters for each of these 10 nights. The observing sequence typically consisted of 300 s integrations on the satellite while in eclipse, alternated with offsets to a nearby sunlit pointing satellite to check positioning and recenter if needed. The total on-target integration time used in the analysis for each observation is given in Table 1.

Table 1. Observing Parameters

Date (UTC)Time (UTC)Target tint (min)Diam. (arcsec)CML a (°W)BG b Avg. Sep. c Notes
1998-11-1505:56–08:55Ganymede401.6310–161.52.2
2018-3-22 d 12:27–14:15Europa500.91350–357380.7Poor weather
2018-6-1507:10–08:54Ganymede201.587–11181.0
2021-5-2013:40–15:05Europa400.87345–3515.41.2
2021-6-812:48–15:17Ganymede85 e 1.55348–3532.12.1
2021-6-2113:05–15:15Europa600.96345–3545.70.9
2021-7-414:13–15:25Callisto251.54349–35013.7
2021-7-1610:12–11:39Europa501.03348–354140.6
2021-9-2609:05–11:03Callisto201.585–91.72.4Poor weather
2021-10-104:55–08:19Ganymede701.707–142.31.3Poor weather

Notes.

a Central meridian longitude, or subobserver longitude. b Sky background level measured near 6300 Å, relative to a measured 21 electrons s−1 arcsec−2 background in the Callisto observations taken on 2021 July 4. c Average angular distance of the target from Jupiter's limb in units of Jupiter radii. d Previously published in de Kleer & Brown (2018). e Some infrared orders are removed during averaging due to additional contamination; total integration times were 70 minutes for 7774 Å and 60 minutes for 8446 Å.

Download table as:  ASCIITypeset image

The HIRES post-upgrade three CCD mosaic covers wavelengths from roughly 5000–9500 Å with our observing setup. The echelle and cross-disperser angles were chosen to ensure that the lines of interest did not fall in the gaps between orders. A slit width of 1farcs722 was used to cover the entire satellite, resulting in a spectral resolution of approximately 24,000. Slit lengths of 7'' and 14'' were used on different nights; a longer slit length increases the amount of sky available for sky subtraction, but results in overlap between the orders at short wavelengths.

To correct for cosmic rays, we used the Laplacian cosmic-ray identification method L.A.Cosmic (van Dokkum 2001) as implemented in the Astropy-affiliated package ccdproc using the method described in McCully et al. (2018). We then bias-subtracted, flat-fielded, and gain-corrected all science images. We used the sharp edges in the flat-field to identify the boundaries of the individual echelle orders, which we used to rectify each order. For the wavelength calibrations, we used thorium-argon arc lamp lines to calculate a wavelength solution for each echelle order by fitting third-degree polynomials. We also corrected for airmass extinction using the wavelength-dependent magnitude attenuation appropriate for the summit of Maunakea derived by Buton et al. (2013). To subtract the background, we calculated a characteristic normalized spatial profile along the slit in the area immediately around the emission line, excluding pixels contained within an aperture covering the target satellite. We then fit this profile and a constant term using ordinary least-squares to subtract the background from each echelle order.

To calibrate each observation from electrons s−1 bin−1 to rayleighs (R), we calculated an expected spectral surface brightness for Jupiter's central meridian using its absolute reflectivity (Woodman et al. 1979) and a solar reference spectrum scaled to the Jupiter-Earth distance at the time of the observations. We then calculated the total number of electrons per second per arcsecond from Jupiter at each auroral wavelength, which provides a direct conversion from electrons to photons. To extract the target surface brightness, we calibrated the data using the apparent angular size of the target disk and the Jupiter flux calibration described above, then fit a Gaussian model and integrated it over a window of ±1 Å around the line center (or each line center for a multiplet emission). We calculated brightnesses from both individual frames and an average of all frames.

The reduced and calibrated spectra in the vicinity of each measured emission line on each date are shown in Figure 1.

Figure 1.

Figure 1. Average spectra centered on each emission line discussed in this paper on each night of observation. All spectra are reported in R Å−1 divided by the angular size of the target. Green spectra are detections with calibrated line strengths; red spectra indicate cloudy nights in which a detection is made, but no quantitative measurement is reported; blue spectra indicate that upper limits are reported; and black spectra are nondetections without reported upper limits due to nonphotometric conditions. Ticks along the horizontal axis shows the location of the center of the emission line(s), and the numbers on the bottom row indicate the wavelength boundaries of each plot in that column. For single emission lines, the boundaries are ±1 Å from the line center. For triplet lines, the boundaries are −1 Å from the shortest wavelength and +1 Å from the longest wavelength. All emission lines arise from atomic oxygen, except for hydrogen at 6563 Å (which is not detected).

Standard image High-resolution image

In order for a detection to be reported instead of an upper limit, the following conditions must be met:

  • 1.  
    The fitted surface brightness indicates a detection at the 2σ level or better,
  • 2.  
    the emission line is centered at the expected Doppler-shifted wavelength given the line-of-sight motion of the target,
  • 3.  
    the emission is localized along the slit, and
  • 4.  
    the spectrum passes a visual inspection to ensure that no false detections are reported due to poor background subtraction or other systematic contributions.

For cloudy or nonphotometric nights, we report whether a detection was made, but do not give quantitative measurements or upper limits due to the uncertainty on the flux calibration. Figure 1 color-codes these cases as described in the caption.

3. Analysis

3.1. Emission Model

All observations were made with the target satellites in Jupiter eclipse, which removes any ambiguity about the relative roles of photoexcitation versus electron excitation. Our model assumes that all auroral emissions are produced by electron-impact direct excitation or dissociative excitation of oxygen-bearing species. The model includes the parent species O, O2, H2O, and CO2. The strength of the observed emission lines is a function of the satellite's atmospheric composition and density, and the density and energy distribution of the incident electrons. Our model is adapted from that described in de Kleer & Brown (2018) and uses experimentally derived emission cross sections. The electron energies and temperatures for Europa's orbital location are taken from Bagenal et al. (2015) and are derived from in situ measurements; we use an electron density of 160 cm−3 with Maxwellian-distributed energies, with 95% of the electron population centered at 20 eV and 5% centered at 250 eV. Note that previous observational studies and modeling papers that interpreted these observations used a four times lower electron density (e.g., Hall et al. 1995, 1998; Roth et al. 2016; Vorburger & Wurz 2021), and column densities from these works need to be scaled down by a factor of 4 for comparison with our results and with other more recent work (de Kleer & Brown 2018, 2019; Roth 2021). The model of Oza et al. (2019) uses an electron density of 70 cm−3, which is intermediate between the two values used in other work.

The electrons at Ganymede's orbit were similarly observed by Voyager to be composed of two populations, with a cold population at 30–60 eV and a hot population at 200 eV, constituting 10% of the total number density of 3–10 cm−3 (Sittler & Strobel 1987). However, Eviatar et al. (2001) demonstrated that the electron populations at Ganymede's orbit are insufficient to explain the brightness of the UV aurora and postulated that the aurora is excited by electrons that are locally accelerated to 100 eV. We follow these and other previous authors in adopting a Maxwellian distribution of electron energies centered at 100 eV, with an electron density of 20 cm−3 (but we note that some modeling work has used a higher density of 70 cm−3; Leblanc et al. 2017). An electron density of 20 cm−3 is consistent with recent in situ measurements from the Ganymede Juno flyby, which took place one day prior to our 2021 June 8 Ganymede eclipse observation (Kurth et al. 2022). Although the electron properties are uncertain and therefore the derived column densities should be viewed with caution, the ratios of the emission lines do not change substantially with the adopted electron properties, and the relative abundances of different species is therefore robust to this uncertainty.

For Callisto, information on the electron energies and densities is even more limited. Voyager measured an electron density of 0.1–1.0 cm−3 at Callisto's orbit (Belcher 1983), while Galileo found 0.01–1.0 cm−3 depending on the plasma sheet distance (Kivelson et al. 2004). The electron temperature estimates from these works range from 35 to 200 eV, with a suprathermal population. We adopt a density of 0.15 cm−3 at 35 eV following Belcher (1983). This is the same electron density as was adopted by Cunningham et al. (2015), but they included a hot population at 1.5 keV, constituting 27% of their total population. The cross-section measurements do not extend to this energy, but we tested adding a hot population at the same ratio at either 250 eV and 1.5 keV (extrapolating the cross sections beyond the measurement energies), and found that the derived column densities change by up to ±30% depending on the electron energies. This is well within the uncertainty in the overall electron densities, therefore we neglect the hot electron population in our modeling.

In addition to extending the model to Callisto, we update the model of de Kleer & Brown (2018) by adding CO2 as a parent molecule. Emission cross sections for electron-impact dissociative excitation of CO2 are not available for several of the UV/optical O i emission lines, and for CO2, our model therefore includes only emission from the O(1S), O(3S), and O(5S) states at 5577, 1304, and 1356 Å, respectively. We adopt the emission cross sections for O(1S) recommended by Itikawa (2002), which are based on the measurements of LeClair & McConkey (1994). We adopt the cross sections for O(3S) from Ajello (1971), but note that their measurements disagree with those of Mumma et al. (1972) even after revised normalization of the latter (Itikawa 2002), and the cross sections are consequently uncertain at the ∼10% level over most electron energies and even higher at the low end of the electron energy distribution. Ajello (1971) also measured the emission from the O(5S) state at 1356 Å. Their emission cross sections are presented relative to those of 1304 Å and are therefore limited by the propagated uncertainties from the O(3S) measurements. Additionally, the authors note that the 1356 Å emission is a lower limit since the O(5S) atom may acquire excess kinetic energy during dissociation beyond the thermal velocity. The limitations on the above measurements and the lack of measurements for other emission lines highlights the need for future laboratory measurements of these cross sections for an interpretation of aurorae at CO2-containing atmospheres throughout the solar system.

Our updated auroral model now also includes the O i emissions at 7774 and 8446 Å from the parent molecules O, O2, and H2O. Both emission features are triplets that are produced by the atomic oxygen transitions (3p 5P)→(3s 5S°) and (3p 3P)→(3s 3S°), respectively. Their radiative cascades to ground produce the well-known 1356 Å and 1304 Å UV lines, and so the cascade contributions into these UV lines is simply equal to the 7774 and 8446 Å brightnesses in Rayleigh units. The cross sections for both emissions following electron impact on O2 are taken from Schulman et al. (1985), on H2O from Beenakker et al. (1974) following the recommendation of Itikawa & Mason (2005), and on O from the excitation cross sections recommended by Laher & Gilmore (1990), which are based on the measurements of Gulcicek et al. (1988) and Gulcicek & Doering (1987).

The method of this emission model relies on some implicit assumptions that warrant a brief discussion. First, the approach assumes that all emissions are optically thin. Lines with the strongest Einstein A coefficients saturate first in their curve of growth, and 1304 Å has the strongest transition probability herein. The brightness of the HST/COS resolved 1304 Å triplet at Ganymede matches the optically thin ratio of 5:3:1 (Roth et al. 2021). Locally, Ganymede's emissions are bright relative to the other satellites, which ensures that opacity effects are negligible overall. A second assumption is that electrons remain warm enough to excite all lines at all altitudes; neutral collisions could cool magnetospheric electrons below the various thresholds to excite the different emissions. Cooling rates in Earth's thermosphere are dominated by vibrational excitation of O2 near 1 × 108 cm−3 (Pavlov & Berrington 1999). This density is expected only very near the surface. A third assumption is negligible collisional quenching of long-lived forbidden transitions. With a lifetime of 134 s (Wiese et al. 1996), O(1D) can be collisionally depopulated before radiating. Thermal O2 gas quenches O(1D) at a rate of 5 × 10−11 cm3 s−1 molecule−1 (Streit et al. 1976), and so the critical density where O2 quenches atoms before radiative decay is 1.5 × 108 cm3–again, a negligible effect over the atmospheric columns.

Table 4 gives the modeled emission rate coefficients for the UV and optical emission lines given the relevant electron energies for Europa, Ganymede, and Callisto. Table 5 gives the corresponding emission line ratios for comparison with observations.

3.2. Retrievals of Atmospheric Composition

We use the emission model described in Section 3.1 to find the best-fit atmospheric column densities for each satellite and date of observation. The model atmosphere is composed of O, O2, and H2O, and the free parameters in the fits are the disk-integrated column densities of each of these constituents. CO2 is considered, but is ultimately not included in the fits because the cross sections are unavailable at all but one of the optical lines. The modeled auroral emissions are fit to the measurements and upper limits of the four oxygen transitions (at 5577, 6300/6364, 7774, and 8446 Å) and Hα (6563 Å). An MCMC algorithm is employed to obtain the joint probability distribution of the three atmospheric constituents and hence obtain the most accurate uncertainties on the free parameters. We use the emcee Python implementation (Foreman-Mackey et al. 2013) of the affine-invariant ensemble sampler for MCMC described by Goodman & Weare (2010). An example output of the MCMC algorithm is shown in Appendix A. In our results, we present our final model with uncertainties; upper limits are presented when a species is found to be present at the <2σ level. In the appendix we also report the single maximum likelihood model for each observation.

The strongest constraints on atmospheric composition come from simultaneous measurements that cover the largest range of emission lines. The atmospheric retrievals on data from a single observation have the advantage that all measurements are simultaneous. However, numerous measurements have also been made of the UV auroral lines (e.g., McGrath et al. 2013; Roth et al. 2016). We therefore also retrieve the atmospheric composition using the average measurements of auroral emissions at each transition, including previous UV measurements.

For the optical transitions, the values in the average fits are averaged over the two dates for each satellite that had good observing conditions and low scattered light (i.e., 1998 November 15 and 2021 June 8 for Ganymede, and 2021 May 20 and 2021 June 21 for Europa; see Table 1). For the UV transitions at Europa, we use values of (40 ± 4) and (80 ± 8) R for the 1304 and 1356 Å emissions, respectively, based on 19 observations summarized in Roth et al. (2016). Roth (2021) also present measurements of the UV emissions from the subjovian hemisphere in and out of eclipse that are about half the average value cited above. While these data are a closer match to the viewing geometry of our observations, Roth (2021) show that the aurora are very similar in and out of eclipse, and we adopt the average values rather than the subjovian values because in the case of Europa, the time variability seems to be dominated by the plasma sheet distance and stochastic variability rather than hemispheric trends. Using the UV eclipse measurements instead of the average UV measurements in our modeling provides a much worse fit.

The situation is different for Ganymede, where the interactions between Ganymede's magnetic field and Jupiter's magnetosphere result in distinct auroral differences between the hemispheres (McGrath et al. 2013). In this case, we use the average of the two available subjovian UV measurements, giving values of (15 ± 2) and (36 ± 2) R for the 1304 and 1356 Å emissions, respectively (Roth et al. 2021). While one of these measurements was made in eclipse and one in sunlight, the values are almost identical. Averaging the leading, trailing, and subjovian hemisphere measurements from that work gives higher values of (23 ± 1) and (46 ± 1) R, which provide a worse fit when the retrievals are run jointly with the optical data.

Ly α emission has also been detected on both satellites; on Europa, a value of 70 ± 26 R was found based on two off-limb averages (Roth et al. 2014a). However, the authors note that these emissions may originate from an extended H cloud around Europa rather than directly from H2O dissociation. At Ganymede, the H2O-sourced Ly α emission level predicted by Roth et al. (2021) of 200 R would be inseparable from spatial variations in the surface reflections (Alday et al. 2017; Roth et al. 2021). We therefore do not include Ly α values in our average fits; including upper limits would not affect the fits because our Hα upper limits provide a much more stringent constraint on H2O as a parent molecule because of the smaller uncertainties due to observing in eclipse.

4. Results and Discussion

We report detections of O i emission at 6300/6364, 5577, and 7774 Å at Europa and Ganymede, and additionally, 8446 Å at Ganymede alone, as well as a detection of 6300/6364 Å emission at Callisto. These constitute the first detections of these lines at any of the icy Galilean satellites, with the exception of the recent measurements of 6300/6364 Å at Europa (de Kleer & Brown 2018, 2019), and also the first detections of 7774 and 8446 Å emission at a planetary body other than Earth. An Earth spectrum covering these emissions obtained with Keck/HIRES from Slanger et al. (2004) is shown in Figure 2 for context. Emission from these lines has also been modeled at Mars and Venus (e.g., Borucki et al. 1996; Gronoff et al. 2008; Bougher et al. 2017), but is not strong enough to have been detected to date. 7774 Å emission was reported in an experiment to detect lightning on Venus (Hansell et al. 1995), but nondetection by a dedicated instrument on board Akatsuki suggests that this was spurious (Lorenz et al. 2019).

Figure 2.

Figure 2. Spectrum covering the 7774 and 8446 Å emissions from Earth's atmosphere, obtained with Keck/HIRES. Figure adapted from Slanger et al. (2004).

Standard image High-resolution image

The 6300/6364 Å emission is seen in every observation of both targets, while the 5577 Å emission is seen on several but not all dates. Cassidy et al. (2008) proposed that electron-excited Na could explain Europa's eclipsed emission in the 2000–10500 Å wavelength range (clear filter) during Cassini's Jupiter flyby. While the foreground Na nebula from Io hinders our ability to measure auroral Na, eclipsed O emissions are more than an order of magnitude brighter than the Na D lines at 5890 and 5896 Å and and thus are a more viable explanation for the broadband Cassini measurement. The 7774 and 8446 Å detections at Ganymede are shown in Figure 3. The detection of 6300 and 6364 Å emission at Callisto is shown in Figure 4. The sole previous measurement of Callisto's oxygen atmosphere was attributed to photodissociation of O2 (Cunningham et al. 2015), and so the emission we report here in shadow constitutes the first definitive detection of Callisto's electron-excited aurora at any wavelength.

Figure 3.

Figure 3. Detections of 7774 and 8446 Å O i emission at Ganymede. For each emission, the lower panel shows a calibrated background-subtracted image of the spectrum that shows that the emission is localized along the slit (vertical axis), while the upper panel shows the spectrum summed across spatial bins containing the localized emission. The vertical ticks between the panels indicate the locations of the Doppler-shifted emission wavelengths for each transition. Each spectrum has been smoothed with a Gaussian kernel to enhance the features relative to the background.

Standard image High-resolution image
Figure 4.

Figure 4. Detections of 6300 and 6364 Å O i emission at Callisto. For each emission, the lower panel shows a calibrated background-subtracted image of the spectrum that shows that the emission is localized along the slit (vertical axis), while the upper panel shows the spectrum summed across spatial bins containing the localized emission. The vertical ticks between the panels indicate the locations of the Doppler-shifted emission wavelengths for each transition. Each spectrum has been smoothed with a Gaussian kernel to enhance the features relative to the background. In the 6300 Å case, the bright points near the top and bottom of the slit just to the right of Callisto are residual from the subtraction of Earth's 6300 Å, which is Doppler-shifted relative to the target.

Standard image High-resolution image

Table 2 provides the disk-averaged surface brightnesses (or 2σ upper limits) of Europa, Ganymede, and Callisto on each observation and for every measured emission line. For poor weather nights, when flux calibration was not reliable, we only report whether each line is detected. Table 3 presents the best-fit atmospheric composition for Europa and Ganymede for each observation, as well as the best-fit composition based on the optical and UV lines averaged across several dates of observation. Note that the exact values of column density are subject to uncertainty due to the poorly constrained density of the electrons exciting the emissions, which vary by a factor of a few depending on the study (e.g., Hall et al. 1995; Bagenal et al. 2015; Bagenal & Dols 2020). However, the relative contribution from different molecules to the emissions is robust to this uncertainty because it is constrained by the ratio of the emission at different wavelengths (given in Table 5). The modeled and measured brightnesses of all emission lines are shown in Figures 5 and 6 and are presented in detail in Appendix B.

Figure 5.

Figure 5. Measured auroral emissions along with the best-fit atmospheric model for Europa for each date of observation and for the average auroral emissions including previous UV measurements (Roth et al. 2016).

Standard image High-resolution image
Figure 6.

Figure 6. Measured auroral emissions along with the best-fit atmospheric model for Ganymede for each date of observation and for the average auroral emissions including previous UV measurements (Roth et al. 2021).

Standard image High-resolution image

Table 2. Measured Auroral Surface Brightnesses and Upper Limits [R] b

 6300 Å6364 Å5577 Å7774 Å8446 Å6563 Å
Satellite/DateO(1D)O(1D)O(1S)O(3p5P)O(3p3P)H α
Ganymede      
1998-11-15139±342.9 ± 1.79.2±1.6<6
2018-6-15127±736 ± 541 ± 4<26<20<30
2021-6-8147.6 ± 1.545.0 ± 1.012.5 ± 0.620 ± 27.3 ± 1.4<1.8
2021-10-1 c detected detected
Average143 ± 244 ± 111 ± 120 ± 27.3 ± 1.4
Europa      
2018-3-22 a , c detected detected
2021-5-20573 ± 14190 ± 737 ± 430 ± 15<22<12
2021-6-21613 ± 12176 ± 642 ± 5<40<80<12
2021-7-16547 ± 14170 ± 835 ± 8<60<60<18
Average593 ± 9183±440±330±15
Callisto      
2021-7-46.1 ± 1.73.3 ± 1.2<2<8<4<3
2021-9-26c

Notes.

a Previously published in de Kleer & Brown (2018). b 2σ upper limits; we require a measurement to be at or above the 2σ level to claim a detection. c On poor weather nights when it was not possible to flux-calibrate the data, we report only whether a given line is detected.

Download table as:  ASCIITypeset image

Table 3. Best-fit Model Atmospheres a

 Column Density (×1014 cm−2)
Satellite/DateO2 OH2O
Europa   
2021-5-204.0 ± 0.1<0.4<0.9
2021-6-213.7 ± 0.21.8 ± 0.8<0.8
2021-7-163.7 ± 0.2<1.2<1.1
Average b 4.1 ± 0.1<0.11.2 ± 0.5
Ganymede   
1998-11-154.5 ± 0.1<3<0.6
2018-6-153.2 ± 0.3<816 ± 3
2021-6-84.8 ± 0.1<0.8<0.5
Average b 4.7 ± 0.1<0.3<0.3
Callisto   
2021-7-440 ± 9

Notes.

a For the following assumed electron parameters, as described and referenced in the text: Europa ne = 160 cm−3 with 95% Maxwellian-distributed in energy about 20 eV and 5% about 250 eV, Ganymede ne = 20 cm−3 Maxwellian-distributed about 100 eV, and Callisto ne = 0.15 cm−3 Maxwellian-distributed about 35 eV. b Average atmospheric compositions are based on fits to the average optical and UV brightnesses across all available dates of observation, as described in Section 3.2.

Download table as:  ASCIITypeset image

For Europa, we find that the atmosphere is dominated by O2 with a column density of 3.7–4.1 × 1014 cm−2. The accurate reproduction of several line ratios via dissociative excitation of O2 effectively rules out the possibility of the atomic atmosphere proposed by Shemansky et al. (2014), and the column density of atomic oxygen is constrained to be <1 × 1013 cm−2. Our models tentatively (2.4σ) indicate the presence of H2O at a column density of 1.2 ± 0.5 × 14 cm−2. Our derived O2 column density is within the range found by previous UV observations; the average brightness at 1356 Å found by Roth et al. (2016) corresponds to an average column density of 3.75 × 1014 cm−2 after converting from the 40 cm−3 electron density used in their calculation to the 160 cm−3 used in ours. Our upper limit on O of <1 − 1013 cm−2 is similarly consistent with the upper limit of 6 × 1012 cm−2 derived from UV eclipse observations (Roth 2021). However, our tentative detection of H2O at 1.2 ± 0.5 × 1014 cm−2 is a factor of 10 below the estimated H2O column density derived from the two UV lines by Roth (2021); this difference is explored in Section 4.1.

For Ganymede, we find that the atmosphere is dominated by O2 with a column density of 3.2–4.8 × 1014 cm−2, and we place an upper limit on O of 3 × 1013 cm−2. This upper limit is consistent with the upper limit of 2 × 1012 cm−2 derived from UV eclipse observations (Roth et al. 2021), although the column density we find for O2 is somewhat higher than the 2.8 × 1014 cm−2 adopted by these authors in their modeling (using the same electron density as in our model: 20 cm−3). We also place an upper limit on H2O of 3 × 1013 cm−2. This is lower by a factor of ∼100 than the H2O abundance inferred by Roth et al. (2021). Water in Ganymede's atmosphere is discussed in more detail in Section 4.1.

These derived O2 column densities roughly match predictions by exosphere modeling work for Europa and Ganymede in eclipse (Leblanc et al. 2017; Oza et al. 2019), although we note that the model predictions are for zenith columns, whereas our observations are disk-integrated and likely dominated by the tangent column at the limb.

The best-fit models for individual nights occasionally find a contribution from O or H2O, which could be physically meaningful or could simply be a reflection of the high uncertainty arising from performing retrievals on data with a low signal-to-noise ratio from individual eclipses. For Ganymede on 2018 June 15, the sky background level was anomalously high (see Table 1), and the inference of H2O should be viewed with skepticism despite the high mathematical significance. For Europa on 2021 June 21, the contribution from O is marginal (at the 2.3σ level), although the data quality is good from this night.

For Callisto, the detection of only one excited O i state (6300/6364 Å), combined with the lack of previous electron-excited UV detections, limits the complexity of models that can be fit to the data. Using a model that assumes a pure O2 atmosphere and an electron density at Callisto of 0.15 cm−3, we find a column density of 4.0 ± 0.9 × 1015 cm−2, which matches the 4.0 × 1015 cm−2 inferred from a UV measurements very well, for which the excitation was attributed to photoelectrons (Cunningham et al. 2015). Callisto is discussed further in Section 4.3.

4.1. H2O-dominated Atmospheres on Europa and Ganymede?

Recently, Roth (2021) and Roth et al. (2021) postulated H2O-dominated atmospheres on the trailing hemispheres on Europa and Ganymede. This claim is based in part on an observed 1356/1304 Å ratio that is lower for both satellites on the trailing hemisphere than on the leading hemisphere. While this difference had previously been attributed to a greater O abundance on the trailing hemisphere (Roth et al. 2016; Molyneux et al. 2018), the new observations included eclipse measurements of the subjovian hemisphere, permitting a stronger upper limit on the column density of O by placing a direct constraint on the contribution to 1304 Å from resonant scattering by O.

Our observations are all made in eclipse and therefore image the subjovian hemisphere. In the case of Europa, our upper limit for O of ∼1 × 1013 cm−2 (see Table 2) is consistent with the upper limit of 6 × 1012 cm−2 derived by Roth (2021). However, our best-fit value for H2O, 1.2 ± 0.5 × 1014 cm−2, is a factor of 10 lower than the disk-averaged H2O abundance of ∼1 × 1015 cm−2 found by Roth (2021) for the sunlit trailing hemisphere. Note that while the column densities presented here depend on the assumed electron properties, we have adopted the same electron properties as used by Roth (2021) and Roth et al. (2021), so the factors by which the derived column densities differ between studies are not sensitive to uncertainty in electron density and are only weakly sensitive to uncertainty in electron temperature. Similarly, while the absolute abundances of species derived from our modeling depends on electron density, the relative abundances do not and are therefore not subject to uncertainty on this parameter.

The presence of H2O in our modeling is only indicated at the 2.4σ level and only after averaging data sets together with the UV; for individual nights, we find only upper limits on H2O in the 0.8–1.1 × 1014 cm−2 range. For all UV and optical oxygen lines except 5577 Å, the emission rates are at least ten times higher for O2 than for H2O, so that these lines are not strongly sensitive to H2O (see Table 4). However, for 5577 Å, the emission rates are comparable between the species. This line, combined with Hα, therefore provide the strongest constraints on water abundance. In particular, the H2O disk-integrated column density from Roth (2021) would produce 70 R of emission at Hα, compared to the 2σ upper limit of 12 R from our Europa observations, in addition to predicting O i line ratios inconsistent with the optical and UV lines.

Table 4. Emission Rate Coefficients for Europa, Ganymede, and Callisto

  Parent Species (×10−10 cm3 s−1)
  EuropaGanymedeCallisto
 Wavelength (Å)OO2 H2OCO2 OO2 H2OCO2 OO2 H2OCO2
O(1S)55774.84.42.5402.4117.6854.46.74.257
O(1D)6300 + 6364331183.17.92009.4231565.2
O(3S)1304375.90.331.150141.33.8375.90.31.1
O(5S)13564.6130.0821.01.1320.333.94.6130.081.0
O(3p3P)84469.83.60.407.29.91.49.83.60.4
O(3p5P)77742.06.80.200.54210.532.06.80.2
Ly α 12169.3339.3
H α 65635.0155.0

Download table as:  ASCIITypeset image

To determine whether an atmosphere containing O, O2, and H2O is preferred over an O2-only atmosphere, we fit the data using both the three-species model and the O2-only model and find that the fit is only marginally better when all three species are included. Given the lack of strong preference for the H2O-containing model and the fact that H2O is only preferred by our average fits (and only at 2.4σ), and not on individual nights, we consider the detection of Europa's H2O at all to be tentative in our data. Moreover, we strongly rule out a water abundance of 2.5 × 1014 cm−2 or higher, or alternatively an H2O/O2 ratio above 0.5, on the eclipsed subjovian hemisphere.

For Ganymede, our observations on the two nights that had clear sky conditions as well as low Jupiter scattered light conditions (see Table 1) place an upper limits on H2O of 3 × 1013 cm−2, or a maximum H2O/O2 ratio of 0.06. In contrast, the modeled ratio for the center trailing hemisphere from the UV data was found to be 12–32, and for the leading hemisphere, it was found to be 2–5 (Roth et al. 2021), with the disk-averaged ratios a factor of a few lower. This H2O abundance contributed 1 R to the disk-averaged 1304 Å emission, but should contribute 23 R and 46 R to the optical emission at 5577 Å and 6563 Å, respectively, both of which are strongly ruled out in our observations.

There are several differences between the optical and UV observations that may be relevant to these different results, in particular, the low H2O/O2 ratios (<1) derived from our optical eclipse data of both satellites, and the much higher ratios of 10–30 determined for the trailing hemispheres in the UV.

First, our observations were made with the targets in shadow, while the UV observations were made with the targets in sunlight. For an H2O atmosphere sustained by sublimation, the column density should be lower in eclipse due to the lower surface temperature. Leblanc et al. (2017) modeled Ganymede's H2O exosphere in eclipse and calculated a collapse by four orders of magnitude from a peak column density in the 1014 to 1016 cm−2 range. However, in this case, the UV auroral brightnesses would also be lower in eclipse than in sunlight, whereas the observed auroral brightnesses of the two satellites are comparable in and out of eclipse (Roth 2021; Roth et al. 2021). A sublimation atmosphere that partially collapses in eclipse therefore cannot reconcile the UV and optical data sets. For Europa, even the O2 atmosphere is modeled to drop by an order of magnitude in eclipse (Oza et al. 2019), which is not consistent with the UV observations either.

The observations also differ in that we target the subjovian hemisphere, while the UV-derived H2O abundances were highest on the trailing hemispheres of both Europa and Ganymede. For Europa, the UV line ratio is also just as low, and hence as inconsistent with pure O2, on the subjovian hemisphere as it is on the trailing hemisphere (Roth 2021). Previous UV data and our optical results together are therefore hard to reconcile with the presence of significant O or H2O on Europa, which seems to require a new explanation for the low 1356/1304 Å ratio at least on the subjovian hemisphere.

For Ganymede, the UV line ratio is higher on the subjovian hemisphere than either leading or trailing (Roth et al. 2021) and is consistent with pure O2. The difference between the UV and optical data for Ganymede can thus be attributed to an H2O atmosphere that is only present on the leading/trailing hemispheres, and not on the subjovian, regardless of whether it is sunlit or eclipsed.

Given the clear differences between hemispheres on the two moons, the upper limit derived for O on the subjovian hemispheres might not be applicable to the trailing hemispheres, and hence, both O and H2O remain candidates to explain the low UV ratio on the trailing hemispheres. If atomic O is primarily a product of electron-impact dissociation of O2, it is similarly predicted to be most abundant at the center of the trailing hemisphere for Europa (Cassidy et al. 2013), so either species would produce a 1356/1304 Å ratio that increases from center to limb of the trailing hemisphere, as observed.

A sputtered H2O atmosphere has also been proposed for both satellites (Johnson et al. 1981). For Ganymede, recent modeled sputtered H2O column densities are in the 1011–1013 cm−2 range, while modeled sublimated column densities are on the order of 1016 cm−2 (Marconi 2007; Leblanc et al. 2017; Vorburger et al. 2022). Sublimated column densities on this scale are ruled out in our data, but the expected sputtered column densities are low enough to be below our detection limits. For Europa, sublimated H2O should be roughly two orders of magnitude lower than for Ganymede based on their surface temperatures, while sputtered H2O is limited to 1013 cm−2 as for Ganymede (Shematovich et al. 2005; Smyth & Marconi 2006; Feistel & Wagner 2007; Plainaki et al. 2013). Both predictions are near our detection limits, so we cannot provide strong constraints on a potential sputtered H2O atmosphere. The expected sputtering production is also longitudinally variable. For Ganymede, Leblanc et al. (2017) showed that the H2O sputtering production should be highest on the leading hemisphere due to the magnetic field geometry, while for Europa, Cassidy et al. (2013) showed that the sputtering rate of all species is highest on the trailing hemisphere.

4.2. Time-variability of Europa's Aurora

While Ganymede's aurora is produced by complex interactions between the incoming plasma and Ganymede's magnetic field, Europa's auroral brightness should be a more straightforward product of the local electron density and the column density of species in its atmosphere. In the UV, the aurora is found to vary by a factor of 5–10 across 71 total observations (Roth et al. 2016); there is a correlation with Europa's distance from the Jovian plasma sheet, although there is significant variability between individual exposures as well as between observations at similar plasma sheet locations. There is no indication that there is a difference in brightness between sunlight and eclipse.

Our individual integrations have a sufficiently high signal-to-noise ratio in the 6300/6364 Å emission for us to analyze the changes in the aurora through each eclipse, during which the plasma sheet distance is changing, but longer-timescale variations in either Europa's atmosphere or in the plasma conditions are not a factor. Figure 7 shows the brightness of the O(1D) transition, the sum of the 6300 and 6364 Å lines, as a function of plasma sheet distance (as defined in Phipps & Bagenal 2021) during each observation. There is an apparent correlation with plasma sheet distance on 2021 July 16 and 2021 June 21, although we note that there is no apparent correlation on 2021 May 20, and even when present, the scatter is significant and the correlation reverses slightly right at the equator. We fit a model of the form $B(z)={B}_{0}{e}^{-{\left(z/H\right)}^{2}}$, where B0 is the brightness as Europa is crossing the centrifugal equator of the plasma sheet (Phipps & Bagenal 2021), z is the absolute-value distance from plasma sheet in RJ , and H is the scale height (Hill et al. 1974; Roth et al. 2016). The free parameters are B0 and H, and the best-fit model is shown in Figure 7. The best fit scale height of 1.67 RJ is in the vicinity of the scale height 1.6 RJ found for the UV aurora (Roth et al. 2016), as well as the plasma torus scale height at Europa's orbit 1.7 RJ from the model of Bagenal & Delamere (2011).

Figure 7.

Figure 7. Brightness of Europa's aurora as a function of plasma sheet distance during each observation. Each observation has been normalized to its average. The dashed curve corresponds to a scale height of 1.67 RJ.

Standard image High-resolution image

The UV and optical observations thus collectively demonstrate that Europa's auroral brightness correlates with magnetic latitude, although there is large scatter about this correlation and other sources of variability may dominate on timescales of minutes to days. We note that Io's aurora also varies by a factor of ∼3 even at a given plasma sheet location (Oliversen et al. 2001; Roth et al. 2014b; Schmidt et al. 2023), and that this is indeed expected based on observed variations in the electron density (Bagenal et al. 2015).

4.3. Callisto

Callisto's atmosphere is the most poorly understood of the Galilean satellites. The presence of an ionosphere at Callisto was indicated by Galileo from the plasma density measured during Callisto flyby (Gurnett et al. 2000) and from radio occultation observations (Kliore et al. 2002). Under the assumption of a predominantly O2 atmosphere (which is supported by modeling work by Liang et al. 2005), the radio occultation measurements were used to infer an O2 column density of 4 × 1016 cm−2 (Kliore et al. 2002). However, the ionosphere was only detected on the sunlit trailing hemisphere, which is the side impacted by Jupiter's corotating plasma, and not on the sunlit leading hemisphere. It was also denser on the sunlit than on the night side of Callisto's terminator, suggesting that photons play a role in producing it. Kliore et al. (2002) hypothesized that the atmosphere is generated by sputtering, but it has been suggested that the presence of the ionosphere should divert plasma around Callisto (Strobel et al. 2002). If this is the case, it may be that sputtering on the night side generates the atmosphere, which becomes ionized by photoelectrons when it moves into sunlight.

The first detection of Callisto's aurora was made with HST/COS of the 1304 and 1356 Å emission lines in sunlight on the leading/Jupiter-facing hemisphere (Cunningham et al. 2015). The observed brightnesses of (3.3 ± 2.8) and (3.2 ± 1.6) R for the two emissions correspond to an O2 column density of 4 × 1015 cm−2 assuming excitation by photoelectrons. The authors argue that photo- rather than magnetospheric electrons are exciting the emissions, on the basis of the fact that photoelectron-excited emission is modeled to be ten times brighter, assuming the magnetospheric electrons penetrate the atmosphere at all instead of being diverted around Callisto (Cunningham et al. 2015). This atmospheric density is five times higher than the CO2 column density of 8 × 1014 cm−2 derived from a Galileo off-limb airglow measurement (Carlson 1999), but ten times lower than the trailing hemisphere O2 density inferred from the ionospheric measurements.

We present the first detection of Callisto's optical aurora from 2021 July 4, specifically, the O i emissions at 6300 and 6364 Å with brightnesses of 6.1 ± 1.7 R and 3.3 ± 1.2 R, respectively. When a purely O2 atmosphere is assumed, these emissions correspond to a column density of (4.0 ± 0.9) × 1015 cm−2, which is identical to the column density inferred from previous UV observations, although our derived column density depends on the assumed electron energies and densities at Callisto, which are poorly constrained at the current time and are likely variable. The UV observations were made in sunlight, and the derived column density was based on the assumption that the aurora was primarily excited by photoelectrons rather than magnetospheric electrons. Our observations were made in eclipse, and the detected emissions must therefore be excited by magnetospheric electrons, removing this ambiguity in interpretation. The match between our column density and that inferred from the past UV data may support the interpretation of the excitation mechanism for the UV aurora. However, the large uncertainty on the electron density, and the unknown variability of Callisto's atmosphere and the electron density at Callisto's orbit, prevent firm conclusions. When we assume a purely O2 atmosphere and excitation by magnetospheric electrons alone, our measured optical emissions would correspond to emission of 0.5 R at 1304 Å and 1.2 R at 1356 Å (see Appendix B). These values are lower than the UV aurora measurements of 3.3 ± 2.8 R and 3.2 ± 1.6 R, which suggests that photoelectrons are indeed needed to excite the observed UV emissions, although the measurements are still consistent within 1–2σ with magnetospheric electron excitation due to the large uncertainties. In addition, the column densities inferred from aurora for the leading/subjovian hemispheres are an order of magnitude lower than the estimate from radio occultation for the trailing hemisphere of 4 × 1016 cm−2 (Kliore et al. 2002; Cunningham et al. 2015), which could be due to enhanced atmospheric generation via sputtering on the trailing hemisphere if confirmed by measurements of both hemispheres using the same technique.

The detection of combined 6300 and 6364 Å emission at 9.4 R along with the upper limit on 5577 Å of 2 R places a lower limit on the O(1D)/O(1S) ratio of 4.75. This rules out H2O as the dominant parent molecule, as it would produce an emission ratio of 1.2 (see Table 5). Atomic O was previously ruled out as the dominant parent species due to an early nondetection of 1304 Å, which is produced in part by resonant scattering (Strobel et al. 2002). CO2 is known to be present in Callisto's atmosphere, but the CO2 column density of 1.1 × 1016 cm−2 that would be required to produce our measured auroral emissions is two orders of magnitude higher than the detected CO2 atmosphere (Carlson 1999). We conclude that molecular O2 is the most plausible parent molecule for the aurora we observe at Callisto.

Table 5. Modeled Emission Ratios for Europa, Ganymede, and Callisto

  Parent Species
  EuropaGanymedeCallisto
RatioWavelengths (Å)OO2 H2OCO2 OO2 H2OCO2 OO2 H2OCO2
Lyα/Hα 1216/65631.92.12.0
O(1D)/Hα (6300 + 6364)/65630.610.610.60
O(1D)/O(1S)(6300 + 6364)/55776.7271.23.3191.25.2231.2
O(1D)/O(5S)(6300 + 6364)/13567.19.2377.36.3297.07.832
O(1D)/O(3S)(6300 + 6364)/13040.89209.30.16147.20.5178
O(1D)/O(3p5P)(6300 + 6364)/7774161715159.618151315
O(1D)/O(3p3P)(6300 + 6364)/84463.3337.61.1206.92277
O(5S)/O(3S)1356/13040.112.20.250.910.0222.20.251.00.072.20.250.97
O(5S)/O(1S)1356/55770.942.90.0330.0260.0462.90.0430.0460.743.00.0380.034

Download table as:  ASCIITypeset image

5. Conclusions

We present new detections of oxygen auroral emissions at the icy Galilean satellites Europa, Ganymede, and Callisto. At Europa, we detect emission at 6300/6364, 5577, and 7774 Å, including the first detection of emissions at the latter two of these wavelengths. At Ganymede, we present the first detections of an aurora at any optical wavelength, including measurements at 6300/6364, 5577, 7774, and 8446 Å. At Callisto, we present the first detection at any wavelength of an aurora that must be excited by magnetospheric electrons, via a detection of the 6300/6364 Å lines. Upper limits are presented for oxygen lines when not detected, and for hydrogen Hα in all observations.

The emissions from Europa and Ganymede are fit with atmospheres that are permitted to contain O, O2, and H2O. The best-fit atmosphere for Europa is predominantly O2 with a column density of 4.1 ± 0.1 × 1014 cm−2, and we find weak evidence for H2O at a column density of 1.2 ± 0.5 × 1014 cm−2. The best-fit atmosphere for Ganymede is exclusively O2 at a column density of 4.7 ± 0.1 × 1014 cm−2; an upper limit of 3 × 1013 cm−2 is placed on both atomic O and H2O. These data place strong constraints on H2O abundance because of the unique sensitivity of the 5577 Å and Hα emissions to the presence of water.

The Callisto aurora indicates an O2 column density of 4.0 ± 0.9 × 1015 cm−2 for our adopted electron properties. This matches the column density inferred from previous UV observations (Cunningham et al. 2015), even though the UV aurora was attributed to photoelectron excitation, whereas we observe emissions in eclipse that must be excited by magnetospheric electrons.

These data collectively demonstrate the power of the optical aurora in providing a collection of emission lines that can clearly differentiate between parent species and provide robust constraints on the atmospheric make-up. The recently launched James Webb Space Telescope will cover the wavelengths of most of the transitions presented here and may provide new insight into not just the strength of the aurorae, but their spatial variations across the satellites.

Support for this work was provided by NASA through grant to program HST-GO-15425 from the Space Telescope Science Institute, which is operated by the Associations of Universities for Research in Astronomy, Incorporated, under NASA contract NAS5-26555. We also gratefully acknowledge support from the NASA Solar System Observations program via grants 80NSSC22K0954 and 80NSSC21K1138. The data presented herein were obtained at the W. M. Keck Observatory, which is operated as a scientific partnership among the California Institute of Technology, the University of California and the National Aeronautics and Space Administration. The Observatory was made possible by the generous financial support of the W. M. Keck Foundation. The authors wish to recognize and acknowledge the very significant cultural role and reverence that the summit of Maunakea has always had within the indigenous Hawaiian community. We are most fortunate to have the opportunity to conduct observations from this mountain.

Facility: Keck/HIRES - .

Appendix A

An example output from the MCMC algorithm employed to determine uncertainties on the best-fit parameters is shown in Figure 8.

Figure 8.

Figure 8. Example MCMC output showing the individual and joint posterior distributions for the column densities of the three atmospheric species, based on the measurements and upper limits for the Ganymede average case. In the fits, the electron energies and densities are fixed at the values described in the text, and the column densities of the three species are the free parameters in the fit.

Standard image High-resolution image

Appendix B: Best-fit Model Properties

Detailed information on the best-fit models presented in the paper is given in Table 6. Our best-fit model parameters with uncertainties are given alongside those of the single maximum likelihood model. The predicted and measured brightness of each auroral line are also given, where the predicted brightnesses correspond to the maximum-likelihood model parameters.

Table 6. Best-fit Models for Europa, Ganymede, and Callisto a

 Column DensityAurora Surface Brightness
 O2 OH2O6300 + 6364 Å 5577 Å7774 Å8446 Å6563 Å1304 Å1356 Å1216 Å
Date(×1014 cm−2)O(1D)O(1S)O(3p5P)O(3p3P)Hα O(3S)O(5S)Lyα
Europa           
2021-5-20           
Result w/Uncertainties4.0 ± 0.1<0.4<0.9
Max-Likelihood Model4.00.000 30.737633144246388311
Measured Brightnesses763 ± 1637 ± 430 ± 15<22<12
2021-6-21           
Result w/Uncertainties3.7 ± 0.21.8 ± 0.8<0.8
Max-Likelihood Model3.71.80.167884046491.2137892.3
Measured Brightnesses789 ± 1342 ± 5<40<80<12
2021-7-16
Result w/Uncertainties3.7 ± 0.2<1.2<1.1
Max-Likelihood Model3.70.370.35717304227356795
Measured Brightnesses717 ± 1635 ± 8<60<60<18
Average b            
Result w/Uncertainties4.1 ± 0.1<0.11.2 ± 0.5
Max-Likelihood Model4.10.021.277734452510408418
Measured Brightnesses776 ± 1040 ± 330 ± 15<22<1240 ± 480 ± 8
Ganymede            
1998-11-15           
Result w/Uncertainties4.5 ± 0.1<3<0.6
Max-Likelihood Model4.50.0010.00418210199013290
Measured Brightnesses181.9 ± 3.49.2 ± 1.6<6
2018-6-15           
Result w/Uncertainties3.2 ± 0.3<816 ± 3
Max-Likelihood Model3.22.716164331615504022108
Measured Brightnesses163 ± 941 ± 4<26<20<30
2021-6-8           
Result w/Uncertainties4.8 ± 0.1<0.8<0.5
Max-Likelihood Model4.80.0050.45193112010114313
Measured Brightnesses192.6 ± 1.812.5 ± 0.620 ± 27.3 ± 1.4<1.8
Average b            
Result w/Uncertainties4.7 ± 0.1<0.3<0.3
Max-Likelihood Model4.70.120.041881020100.115300.3
Measured Brightnesses187 ± 211 ± 120 ± 27.3 ± 1.4<215 ± 236 ± 2
Callisto            
2021-7-4           
Result w/Uncertainties40 ± 99.30.40.70.300.51.20.3
Measured Brightnesses9.4 ± 2.1<2<8<4<3
 

Notes.

a Our reported best-fit model with uncertainties is given under the heading "Result w/Uncertainties." However, because the column densities are often upper limits, we also report the single model that maximizes the likelihood function along with the auroral surface brightnesses of that model. b Average auroral brightnesses are averaged over values from this work and from the literature (Roth et al. 2016; de Kleer & Brown 2018, 2019).

Download table as:  ASCIITypeset image

Please wait… references are loading.
10.3847/PSJ/acb53c