67 results on '"A. P. Ingersoll"'
Search Results
2. How the martian residual south polar cap develops quasi-circular and heart-shaped pits, troughs, and moats
- Author
-
James W. Head, Caleb I. Fassett, Bethany L. Ehlmann, P. B. Buhler, and Andrew P. Ingersoll
- Subjects
Martian ,geography ,geography.geographical_feature_category ,010504 meteorology & atmospheric sciences ,Meteorology ,Landform ,Astronomy and Astrophysics ,Fault scarp ,01 natural sciences ,Space and Planetary Science ,0103 physical sciences ,Sublimation (phase transition) ,Ice caps ,Polar cap ,010303 astronomy & astrophysics ,Geomorphology ,Geology ,0105 earth and related environmental sciences - Abstract
The martian Residual South Polar Cap (RSPC) is a 1-10 m thick deposit of permanent CO_2 ice perched on the much larger H_2O ice cap. The CO_2 ice is dissected into mesas by erosional landforms that can be broadly classified as (i) quasi-circular pits, (ii) heart-shaped pits, (iii) linear troughs, and (iv) moats. We use HiRISE (25-50 cm/px) images taken at a cadence of days to months to track meter-scale changes in the RSPC in order to investigate the mechanisms that lead to the development of these four distinct morphologies. For the first time, we report the development of dark fans on the sides of the CO_2 mesas and the fracturing and deterioration of the initially smooth upper surface of CO_2 mesas. We interpret these features as indicating the sublimation and subsequent escape of CO_2 from the interiors of mesas, which undermines structural support of mesa tops, causing them to collapse. The collapse of mesa tops, along with uneven deposition of CO_2 ice, creates steep scarps that erode during the summer due to preferential sunlight absorption. During the winter, CO_2 deposition acts to smooth topography, creating gently sloping ramps. We propose that the interplay between the steep scarps and gentle slopes leads to either quasi-circular pits, heart-shaped pits, linear troughs, or moats, depending on local conditions.
- Published
- 2017
- Full Text
- View/download PDF
3. Controlled boiling on Enceladus. 2. Model of the liquid-filled cracks
- Author
-
Miki Nakajima and Andrew P. Ingersoll
- Subjects
Materials science ,010504 meteorology & atmospheric sciences ,Meteorology ,Vapor pressure ,Astronomy and Astrophysics ,Mechanics ,01 natural sciences ,Plume ,Neutral buoyancy ,Space and Planetary Science ,Boiling ,Saturn ,0103 physical sciences ,Total pressure ,Enceladus ,010303 astronomy & astrophysics ,Water vapor ,0105 earth and related environmental sciences - Abstract
Controlled boiling will occur on Enceladus whenever a long, narrow conduit connects liquid water to the vacuum of space. In a companion paper we focus on the upward flow of the vapor and show how it controls the evaporation rate through backpressure, which arises from friction on the walls. In this paper we focus on the liquid and show how it flows through the conduit up to its level of neutral buoyancy. For an ice shell 20 km thick, the liquid water interface could be 2 km below the surface. We find that the evaporating surface can be narrow. There is no need for a large vapor chamber that acts as a plume source. Freezing on the icy walls and the evaporating surface is avoided if the crack width averaged over the length of the tiger stripes is greater than 1 m and the salinity of the liquid is greater than 20 g kg^(−1). Controlled boiling plays a crucial role in our model, which makes it different from earlier published models. The liquids on Enceladus are boiling because there is no overburden pressure—the saturation vapor pressure is equal to the total pressure. Salinity plays a crucial role in preventing freezing, and we argue that the subsurface oceans of icy satellites can have water vapor plumes only if their salinities are greater than about 20 g kg^(−1).
- Published
- 2016
- Full Text
- View/download PDF
4. Controlled boiling on Enceladus. 1. Model of the vapor-driven jets
- Author
-
Miki Nakajima and Andrew P. Ingersoll
- Subjects
Materials science ,010504 meteorology & atmospheric sciences ,Meteorology ,Astronomy and Astrophysics ,Mechanics ,01 natural sciences ,Plume ,Volumetric flow rate ,Space and Planetary Science ,Boiling ,Latent heat ,0103 physical sciences ,Fluid dynamics ,Mass flow rate ,Enceladus ,010303 astronomy & astrophysics ,Water vapor ,0105 earth and related environmental sciences - Abstract
Plumes of water vapor and ice particles have been observed from the so-called tiger stripes at the south polar terrain (SPT) of Saturn’s satellite, Enceladus. The observed high salinity (∼0.5–2%) of the ice particles in the plumes may indicate that the plumes originate from a subsurface liquid ocean. Additionally, the SPT is the source of strong infrared radiation (∼4.2 GW), which is especially intense near (within tens of meters) the tiger stripes. This could indicate that the radiation is associated with plume activity, but the connection remains unclear. Here we investigate the constraints that plume observations place on the widths of the cracks, the depth to the liquid-vapor interface, and the mechanisms controlling plume variability. We solve the fluid dynamics of the flow in the crack and the interaction between the flow and ice walls assuming that the flows of water vapor and ice particles originate from a few kilometers deep liquid ocean. For a crack with a uniform width, we find that our model could explain the observed vapor mass flow rate of the plumes when the crack width is 0.05–0.075 m. A wider crack is not favorable because it would produce a higher vapor mass flow rate than the observed value, but it may be allowed if there are some flows that do not reach the surface of Enceladus either due to condensation on the icy walls or the tortuosity of the crack. The observed heat flow can be explained if the total crack length is approximately 1.7 × 500 km. A tapering crack (a crack which is ∼1 m wide at the bottom of the flow and sharply becomes 0.05–0.075 m at shallower depths) can also explain the observed vapor mass flow rate and heat flow. Widths of 1 m or more are necessary to avoid freezing at the liquid-vapor interface, as shown in our paired paper (Ingersoll and Nakajima [2016] Icarus). The observed intense heat flow along the tiger stripes can be explained by the latent heat release due to vapor condensation onto the ice walls near the surface. The resulting buildup of ice causes the vents to seal themselves on time scales less than a year. We also find that the ice to vapor ratio of the plumes is sensitive to the ice mass fraction at the bottom of the flow (liquid–vapor interface). We find that the total mass flow rate of the plumes becomes larger when the crack width is larger, which is consistent with the observation that the flow rate increases near the orbital apocenter, where the crack is expected to be widest.
- Published
- 2016
- Full Text
- View/download PDF
5. Spatial distribution of ice blocks on Enceladus and implications for their origin and emplacement
- Author
-
Bernd Giese, Shawn P. Ewald, Andrew P. Ingersoll, Hilary R. Martens, and Paul Helfenstein
- Subjects
surface features ,Astronomy and Astrophysics ,Mass wasting ,Pressure ridge ,Spatial distribution ,Enceladus ,Tectonics ,Paleontology ,Impact crater ,ice blocks ,Space and Planetary Science ,Lithosphere ,Southern Hemisphere ,Geology - Abstract
We have mapped the locations of over 100,000 ice blocks across the south polar region of Saturn’s moon Enceladus, thus generating the first quantitative estimates of ice-block number density distribution in relation to major geological features. Ice blocks were manually identified and mapped from twenty of the highest resolution (4–25 m per pixel) Cassini Imaging Science Subsystem (ISS) narrow-angle images using ArcGIS software. The 10–100 m-diameter positive-relief features are marginally visible at the resolution of the images, making ice-block identifications difficult but not impossible. Our preliminary results reveal that ice blocks in the southern hemisphere are systematically most concentrated within the geologically active South Polar Terrain (SPT) and exhibit peak concentrations within 20 km of the tiger-stripe fractures as well as close to the south pole. We find that ice blocks are concentrated just as heavily between tiger-stripe fractures as on the directly adjacent margins; although significant local fluctuations in ice-block number density do occur, we observe no clear pattern with respect to the tiger stripes or jet sources. We examine possible roles of several mechanisms for ice-block origin, emplacement, and evolution: impact cratering, ejection from fissures during cryovolcanic eruptions, tectonic disruption of lithospheric ice, mass wasting, seismic disturbance, and vapor condensation around icy fumeroles. We conclude that impact cratering as well as mass wasting, perhaps triggered by seismic events, cannot account for a majority of ice-block features within the inner SPT. The pervasiveness of fracturing at many size scales, the ubiquity of ice blocks in the inner SPT, as well as the occurrence of linear block arrangements that parallel through-cutting crack networks along the flanks of tiger stripes indicate that tectonic deformation is an important source of blocky-ice features in the SPT. Ejection during catastrophic cryovolcanic eruptions and condensation around surface vents, however, cannot be ruled out. Further, sublimation processes likely erode and disaggregate ice blocks from solid exposures of ice, especially near the warm tiger-stripe fractures. The relative paucity of blocks beyond the bounds of the SPT, particularly on stratigraphically old cratered terrains, may be explained in part by mantling of the surface by fine particulate ice grains that accumulate over time.
- Published
- 2015
- Full Text
- View/download PDF
6. Cassini ISS observation of Saturn’s String of Pearls
- Author
-
Shawn P. Ewald, Kunio M. Sayanagi, Gabriel D. Muro, Ulyana A. Dyudina, and Andrew P. Ingersoll
- Subjects
Physics ,ICARUS ,Planetary science ,Space and Planetary Science ,Gas giant ,Turbulent wake ,Astronomy ,Astronomy and Astrophysics ,Storm ,Longitude ,Merge (version control) ,Latitude - Abstract
We present the dynamics of the String of Pearls (SoPs) feature observed by the Cassini spacecraft’s Imaging Science Subsystem (ISS) camera between 2007 and 2010. The SoPs was originally discovered in the 5 μm images captured by Cassini VIMS instrument, where it appeared as a chain of infrared-bright spots (Momary, T.W., et al. [2006]. The Zoology of Saturn: The Bizarre Features Unveiled by the 5 Micron Eyes of Cassini/VIMS. AAS/Division for Planetary Sciences Meeting Abstracts 38, 499). Using ISS images of Saturn, we found a chain of 23–26 dark spots at 33.2°N planetocentric latitude with characteristics that are consistent with those of SoPs. Our measurements imply that the feature propagated at −2.26 ± 0.02° day^−1 in longitude (−22.27 ± 0.2 m s^−1, negative values denote westward) during the observed period that spans three Earth years. Our measurements imply that the SoPs is a chain of cyclones, which we infer from the motion of clouds on the periphery of the individual pearls. We tracked the motion of 26 pearls for 6 months in 2008 and noted a few pearls appearing and disappearing, all near the east–west termini of the SoPs feature. During this period, a few of the pearls, varying between 6 and 10, harbored a small circular cloud at the center, which we call the central peaks. In general, a group of vortices with the same sign of vorticity tend to merge; however, our measurements did not detect merger of pearls. The interest in the feature was heightened when the latest planet-encircling storm erupted from the SoPs on December 5, 2010 (Sayanagi, K.M., Dyudina, U.A., Ewald, S.P., Fischer, G., Ingersoll, A.P., Kurth, W.S., Muro, G.D., Porco, C.C., West, R.A. [2013]. Icarus 223, 460–478). The storm severely disrupted the region; the SoPs was last seen on December 24, 2010 in the turbulent wake of the storm, and has not reappeared as of August 2013.
- Published
- 2014
- Full Text
- View/download PDF
7. Saturn’s thermal emission at 2.2-cm wavelength as imaged by the Cassini RADAR radiometer
- Author
-
Michael Allison, Kevin H. Baines, Andrew P. Ingersoll, K. Kelleher, Y. Anderson, A. L. Laraia, Fabiano Oyafuso, Samuel Gulkis, Scott Edgington, and M. A. Janssen
- Subjects
Physics ,Radiometer ,Atmospheric models ,Astronomy ,Astronomy and Astrophysics ,Latitude ,Atmosphere ,Wavelength ,Space and Planetary Science ,Brightness temperature ,Saturn ,Physics::Space Physics ,Radiative transfer ,Astrophysics::Earth and Planetary Astrophysics ,Physics::Atmospheric and Oceanic Physics ,Astrophysics::Galaxy Astrophysics - Abstract
We present well-calibrated, high-resolution maps of Saturn’s thermal emission at 2.2-cm wavelength obtained by the Cassini RADAR radiometer through the Prime and Equinox Cassini missions, a period covering approximately 6 years. The absolute brightness temperature calibration of 2% achieved is more than twice better than for all previous microwave observations reported for Saturn, and the spatial resolution and sensitivity achieved each represent nearly an order of magnitude improvement. The brightness temperature of Saturn in the microwave region depends on the distribution of ammonia, which our radiative transfer modeling shows is the only significant source of absorption in Saturn’s atmosphere at 2.2-cm wavelength. At this wavelength the thermal emission comes from just below and within the ammonia cloud-forming region, and yields information about atmospheric circulations and ammonia cloud-forming processes. The maps are presented as residuals compared to a fully saturated model atmosphere in hydrostatic equilibrium. Bright regions in these maps are readily interpreted as due to depletion of ammonia vapor in, and, for very bright regions, below the ammonia saturation region. Features seen include the following: a narrow equatorial band near full saturation surrounded by bands out to about 10° planetographic latitude that demonstrate highly variable ammonia depletion in longitude; narrow bands of depletion at −35° latitude; occasional large oval features with depleted ammonia around −45° latitude; and the 2010–2011 storm, with extensive saturated and depleted areas as it stretched halfway around the planet in the northern hemisphere. Comparison of the maps over time indicates a high degree of stability outside a few latitudes that contain active regions.
- Published
- 2013
- Full Text
- View/download PDF
8. Analysis of Saturn’s thermal emission at 2.2-cm wavelength: Spatial distribution of ammonia vapor
- Author
-
A. L. Laraia, Michael Janssen, Fabiano Oyafuso, Andrew P. Ingersoll, Michael Allison, and Samuel Gulkis
- Subjects
Brightness ,Materials science ,Astronomy and Astrophysics ,Astrophysics::Cosmology and Extragalactic Astrophysics ,Atmospheric model ,Atmospheric sciences ,Computational physics ,Troposphere ,Atmosphere ,Space and Planetary Science ,Saturn ,Brightness temperature ,Radiative transfer ,Mixing ratio ,Astrophysics::Solar and Stellar Astrophysics ,Astrophysics::Earth and Planetary Astrophysics ,Physics::Atmospheric and Oceanic Physics ,Astrophysics::Galaxy Astrophysics - Abstract
This work focuses on determining the latitudinal structure of ammonia vapor in Saturn's cloud layer near 1.5 bars using the brightness temperature maps derived from the Cassini RADAR (Elachi et al., 2004) instrument, which works in a passive mode to measure thermal emission from Saturn at 2.2-cm wavelength. We perform an analysis of five brightness temperature maps that span epochs from 2005 to 2011, which are presented in a companion paper by Janssen et al. (2013a, this issue). The brightness temperature maps are representative of the spatial distribution of ammonia vapor, since ammonia gas is the only effective opacity source in Saturn's atmosphere at 2.2-cm wavelength. Relatively high brightness temperatures indicate relatively low ammonia relative humidity (RH), and vice versa. We compare the observed brightness temperatures to brightness temperatures computed using the Juno atmospheric microwave radiative transfer (JAMRT) program which includes both the means to calculate a tropospheric atmosphere model for Saturn and the means to carry out radiative transfer calculations at microwave frequencies. The reference atmosphere to which we compare has a 3x solar deep mixing ratio of ammonia (we use 1.352x10(exp -4) for the solar mixing ratio of ammonia vapor relative to H2; see Atreya, 2010) and is fully saturated above its cloud base. The maps are comprised of residual brightness temperatures-observed brightness temperature minus the model brightness temperature of the saturated atmosphere.
- Published
- 2013
- Full Text
- View/download PDF
9. Dynamics of Saturn’s great storm of 2010–2011 from Cassini ISS and RPWS
- Author
-
Andrew P. Ingersoll, William S. Kurth, Georg Fischer, Robert A. West, Kunio M. Sayanagi, Carolyn C. Porco, Ulyana A. Dyudina, Shawn P. Ewald, and Gabriel D. Muro
- Subjects
Earth and Planetary Astrophysics (astro-ph.EP) ,Solar System ,Astrophysics::High Energy Astrophysical Phenomena ,FOS: Physical sciences ,Astronomy and Astrophysics ,Storm ,Atmospheric sciences ,Lightning ,Latitude ,Vortex ,Jupiter ,Space and Planetary Science ,Anticyclone ,Saturn ,Climatology ,Physics::Space Physics ,Astrophysics::Solar and Stellar Astrophysics ,Astrophysics::Earth and Planetary Astrophysics ,Physics::Atmospheric and Oceanic Physics ,Geology ,Astrophysics - Earth and Planetary Astrophysics - Abstract
Saturn's quasi-periodic planet-encircling storms are the largest convecting outbursts in the Solar System. The last eruption was in 1990. A new eruption started in December 2010 and presented the first-ever opportunity to observe such episodic storms from a spacecraft in orbit around Saturn. Here, we analyze images acquired with the Cassini Imaging Science Subsystem (ISS), which captured the storm's birth, evolution and demise. In studying the end of the convective activity, we also analyze the Saturn Electrostatic Discharge (SED) signals detected by the Radio and Plasma Wave Science (RPWS) instrument. [...], 17 figures
- Published
- 2013
- Full Text
- View/download PDF
10. Emergence of polar-jet polygons from jet instabilities in a Saturn model
- Author
-
Kunio M. Sayanagi, Raul Morales-Juberias, Timothy E. Dowling, and Andrew P. Ingersoll
- Subjects
Physics ,Jet (fluid) ,Astrophysics::High Energy Astrophysical Phenomena ,Astronomy ,Astronomy and Astrophysics ,Geometry ,Jet stream ,Wind speed ,Kármán vortex street ,Vortex ,Astrophysical jet ,Space and Planetary Science ,Saturn ,Physics::Space Physics ,Astrophysics::Earth and Planetary Astrophysics ,Physics::Atmospheric and Oceanic Physics ,Saturn's hexagon - Abstract
Voyager flybys of Saturn in 1980–1981 revealed a circumpolar wave at ≈78° north planetographic latitude. The feature had a dominant wavenumber 6 mode, and has been termed the Hexagon from its geometric appearance in polar-projected mosaics. It was also noted for being stationary with respect to Saturn’s Kilometric Radiation (SKR) rotation rate. The Hexagon has persisted for over 30 years since the Voyager observations until now. It has been observed from ground based telescopes, Hubble Space Telescope and multiple instruments onboard Cassini in orbit around Saturn. Measurements of cloud motions in the region reveal the presence of a jet stream whose path closely follows the Hexagon’s outline. Why the jet stream takes the characteristic six-sided shape and how it is stably maintained across multiple saturnian seasons are yet to be explained. We present numerical simulations of the 78.3°N jet using the Explicit Planetary Isentropic-Coordinate (EPIC) model and demonstrate that a stable hexagonal structure can emerge without forcing when dynamic instabilities in the zonal jet nonlinearly equilibrate. For a given amplitude of the jet, the dominant zonal wavenumber is most strongly dependent on the peak curvature of the jet, i.e., the second north–south spatial derivative of the zonal wind profile at the center of the jet. The stable polygonal shape of the jet in our simulations is formed by a vortex street with cyclonic and anticyclonic vortices lining up towards the polar and equatorial side of the jet, respectively. Our result is analogous to laboratory experiments of fluid motions in rotating tanks that develop polygonal flows out of vortex streets. However, our results also show that a vortex street model of the Hexagon cannot reproduce the observed propagation speed unless the zonal jet’s speed is modified beyond the uncertainties in the observed zonal wind speed, which suggests that a vortex street model of the Hexagon and the observed zonal wind profile may not be mutually compatible.
- Published
- 2011
- Full Text
- View/download PDF
11. Subsurface heat transfer on Enceladus: Conditions under which melting occurs
- Author
-
Andrew P. Ingersoll and Alexey A. Pankine
- Subjects
Materials science ,Space and Planetary Science ,Vapor pressure ,Latent heat ,Heat spreader ,Heat transfer ,Condensation ,Thermal contact ,Thermodynamics ,Astronomy and Astrophysics ,Mechanics ,Diffusion (business) ,Thermal conduction - Abstract
Given the heat that is reaching the surface from the interior of Enceladus, we ask whether liquid water is likely and at what depth it might occur. The heat may be carried by thermal conduction through the solid ice, by the vapor as it diffuses through a porous matrix, or by the vapor flowing upward through open cracks. The vapor carries latent heat, which it acquires when ice or liquid evaporates. As the vapor nears the surface it may condense onto the cold ice, or it may exit the vent without condensing, carrying its latent heat with it. The ice at the surface loses its heat by infrared radiation. An important physical principle, which has been overlooked so far, is that the partial pressure of the vapor in the pores and in the open cracks is nearly equal to the saturation vapor pressure of the ice around it. This severely limits the ability of ice to deliver the observed heat to the surface without melting at depth. Another principle is that viscosity limits the speed of the flow, both the diffusive flow in the matrix and the hydrodynamic flow in open cracks. We present hydrodynamic models that take these effects into account. We find that there is no simple answer to the question of whether the ice melts or not. Vapor diffusion in a porous matrix can deliver the heat to the surface without melting if the particle size is greater than ~1 cm and the porosity is greater than ~0.1, in other words, if the matrix is a rubble pile. Whether such an open matrix can exist under its own hydrostatic load is unclear. Flow in open cracks can deliver the heat without melting if the width of the crack is greater than ~10 cm, but the heat source must be in contact with the crack. Frictional heating on the walls due to tidal stresses is one such possibility. The lifetime of the crack is a puzzle, since condensation on the walls in the upper few meters could seal the crack off in a year, and it takes many years for the heat source to warm the walls if the crack extends down to km depths. The 10:1 ratio of radiated heat to latent heat carried with the vapor is another puzzle. The models tend to give a lower ratio. The resolution might be that each tiger stripe has multiple cracks that share the heat, which tends to lower the ratio. The main conclusion is that melting depends on the size of the pores and the width of the cracks, and these are unknown at present.
- Published
- 2010
- Full Text
- View/download PDF
12. Lightning storms on Saturn observed by Cassini ISS and RPWS during 2004–2006
- Author
-
Ulyana A. Dyudina, Carolyn C. Porco, William S. Kurth, Georg Fischer, Joseph Ferrier, Shawn P. Ewald, Anthony D. Del Genio, J. Barbara, Michael D. Desch, and Andrew P. Ingersoll
- Subjects
Physics ,Space and Planetary Science ,Planet ,Thunderstorm ,Astronomy ,Astronomy and Astrophysics ,Storm ,Orbit insertion ,Latitude - Abstract
We report on Cassini Imaging Science Subsystem (ISS) data correlated with Radio and Plasma Wave Science (RPWS) observations, which indicate lightning on Saturn. A rare bright cloud erupt at ∼35° South planetocentric latitude when radio emissions (Saturn Electrostatic Discharges, or SEDs) occur. The cloud consisting of few consecutive eruptions typically lasts for several weeks, and then both the cloud and the SEDs disappear. They may reappear again after several months or may stay inactive for a year. Possibly, all the clouds are produced by the same atmospheric disturbance which drifts West at 0.45 °/day. As of March 2007, four such correlated visible and radio storms have been observed since Cassini Saturn Orbit Insertion (July 2004). In all four cases the SEDs are periodic with roughly Saturn's rotation rate (h^(10)m^(39)), and show correlated phase relative to the times when the clouds are seen on the spacecraft-facing side of the planet, as had been shown for the 2004 storms in [Porco, C.C., and 34 colleagues, 2005. Science 307, 1243–1247]. The 2000-km-scale storm clouds erupt to unusually high altitudes and then slowly fade at high altitudes and spread at low altitudes. The onset time of individual eruptions is less than a day during which time the SEDs reach their maximum rates. This suggests vigorous atmospheric updrafts accompanied by strong precipitation and lightning. Unlike lightning on Earth and Jupiter, where considerable lightning activity is known to exist, only one latitude on Saturn has produced lightning strong enough to be detected during the two and a half years of Cassini observations. This may partly be a detection issue.
- Published
- 2007
- Full Text
- View/download PDF
13. Analysis of a giant lightning storm on Saturn
- Author
-
M. L. Kaiser, D. A. Gurnett, Andrew P. Ingersoll, P. Zarka, Alain Lecacheux, Ulyana A. Dyudina, William S. Kurth, Georg Fischer, Department of Physics and Astronomy, Iowa State University, Geological and Planetary Sciences, California Institute of Technology, Pasadena, NASA/Goddard Space Flight Center (NASA/GSFC), Observatoire de Paris, Université Paris sciences et lettres (PSL), Laboratoire d'études spatiales et d'instrumentation en astrophysique (LESIA), Université Pierre et Marie Curie - Paris 6 (UPMC)-Institut national des sciences de l'Univers (INSU - CNRS)-Observatoire de Paris, Université Paris sciences et lettres (PSL)-Université Paris sciences et lettres (PSL)-Université Paris Diderot - Paris 7 (UPD7)-Centre National de la Recherche Scientifique (CNRS), Physique des plasmas, and Université Paris sciences et lettres (PSL)-Université Paris sciences et lettres (PSL)-Université Paris Diderot - Paris 7 (UPD7)-Centre National de la Recherche Scientifique (CNRS)-Université Pierre et Marie Curie - Paris 6 (UPMC)-Institut national des sciences de l'Univers (INSU - CNRS)-Observatoire de Paris
- Subjects
Astrophysics::High Energy Astrophysical Phenomena ,Ionospheric electron density ,Astronomy ,Astronomy and Astrophysics ,Storm ,Astrophysics::Cosmology and Extragalactic Astrophysics ,Lightning ,Latitude ,Atmosphere ,Space and Planetary Science ,Saturn ,Physics::Space Physics ,Astrophysics::Solar and Stellar Astrophysics ,[PHYS.ASTR]Physics [physics]/Astrophysics [astro-ph] ,Astrophysics::Galaxy Astrophysics ,Geology - Abstract
International audience; On January 23, 2006, the Cassini/RPWS (Radio and Plasma Wave Science) instrument detected a massive outbreak of SEDs (Saturn Electrostatic Discharges). The following SED storm lasted for about one month and consisted of 71 consecutive episodes. It exceeded all other previous SED observations by Cassini as well as by the Voyagers with regard to number and rate of detected events. At the same time astronomers at the Earth as well as Cassini/ISS (Imaging Science Subsystem) detected a distinctive bright atmospheric cloud feature at a latitude of 35° South, strongly confirming the current interpretation of SEDs being the radio signatures of lightning flashes in Saturn's atmosphere. In this paper we will analyze the main physical properties of this SED storm and of a single small SED storm from 2005. The giant SED storm of 2006 had maximum burst rates of 1 SED every 2 s, its episodes lasted for 5.5 h on average, and the episode's periodicity of about 10.66 h exactly matched the period of the ISS observed cloud feature. Using the low frequency cutoff of SED episodes we determined an ionospheric electron density around 10 cm for the dawn side of Saturn.
- Published
- 2007
- Full Text
- View/download PDF
14. Saturn eddy momentum fluxes and convection: First estimates from Cassini images
- Author
-
Robert A. West, J. Barbara, Carolyn C. Porco, Joseph Ferrier, Ashwin R. Vasavada, Joseph N. Spitale, Andrew P. Ingersoll, and Anthony D. Del Genio
- Subjects
Convection ,Physics ,Solar System ,Gas giant ,Momentum transfer ,Astronomy and Astrophysics ,Geophysics ,Kinetic energy ,Atmospheric sciences ,Physics::Fluid Dynamics ,Atmosphere ,Eddy ,Space and Planetary Science ,Physics::Space Physics ,Convective storm detection ,Astrophysics::Earth and Planetary Astrophysics ,Physics::Atmospheric and Oceanic Physics - Abstract
We apply an automated cloud feature tracking algorithm to estimate eddy momentum fluxes in Saturn's southern hemisphere from Cassini Imaging Science Subsystem near-infrared continuum image sequences. Voyager Saturn manually tracked images had suggested no conversion of eddy to mean flow kinetic energy, but this was based on a small sample of ∼ 10 m 2 s −2 and a clear positive correlation between eddy momentum fluxes and meridional shear of the mean zonal wind, implying that eddies supply momentum to eastward jets and remove momentum from westward jets at a rate ∼ 5 × 10 −6 m s −2 . The behavior we observe is similar to that seen on Jupiter, though with smaller eddy-mean kinetic energy conversion rates per unit mass of atmosphere ( 3.3 × 10 −5 m 2 s −3 ). We also use the appearance and rapid evolution of small bright features at continuum wavelengths, in combination with evidence from weak methane band images where possible, to diagnose the occurrence of moist convective storms on Saturn. Areal expansion rates imply updraft speeds of ∼ 1 m s −1 over the convective anvil cloud area. As on Jupiter, convection preferentially occurs in cyclonic shear regions on Saturn, but unlike Jupiter, convection is also observed in eastward jet regions. With one possible exception, the large eddy fluxes seen in the cyclonic shear latitudes do not seem to be associated with convective events.
- Published
- 2007
- Full Text
- View/download PDF
15. Interaction between eddies and mean flow in Jupiter's atmosphere: Analysis of Cassini imaging data
- Author
-
Colette Salyk, Anthony D. Del Genio, Jean J. Lorre, Ashwin R. Vasavada, and Andrew P. Ingersoll
- Subjects
Physics ,Jupiter ,Momentum ,Atmosphere ,Turbulent diffusion ,Heat flux ,Eddy ,Space and Planetary Science ,Atmosphere of Jupiter ,Zonal flow ,Astronomy and Astrophysics ,Astrophysics ,Atmospheric sciences - Abstract
Beebe et al. [Beebe, R.F., et al., 1980. Geophys. Res. Lett. 17, 1–4] and Ingersoll et al. [Ingersoll, A.P., et al., 1981. J. Geophys. Res. 86, 8733–8743] used images from Voyagers 1 and 2 to analyze the interaction between zonal winds and eddies in Jupiter's atmosphere. They reported a high positive correlation between Jupiter's eddy momentum flux, pu'v', and the variation of zonal velocity with latitude, du/dy. This correlation implied a surprisingly high rate of conversion of energy from eddies to zonal flow: ~1.5-3.0 Wm^(-2), a value more than 10% of Jupiter's thermal flux emission. However, Sromovsky et al. [Sromovsky, L.A., et al., 1982. J. Atmos. Sci. 39, 1413–1432] argued that possible biases in the analysis could have caused an artificially high correlation. In addition, significant differences in the derived eddy flux between datasets put into question the robustness of any one result. We return to this long-standing puzzle using images of Jupiter from the Cassini flyby of December 2000. Our method is similar to previous analyses, but utilizes an automatic feature tracker instead of the human eye. The number of velocity vectors used in this analysis is over 200,000, compared to the 14,000 vectors used by Ingersoll et al. We also find a positive correlation between u'v' and du/dy and derive a global average power per unit mass, u'v'du/dy, ranging from (7.1-12.3 x 10^(-5) Wkg^(-1). Utilizing Ingersoll et al.'s estimate of the mass per unit area involved in the transport, this would imply a rate of energy conversion of ~0.7-1.2 Wm^(-2). We discuss the implications of this result and employ several tests to demonstrate its robustness.
- Published
- 2006
- Full Text
- View/download PDF
16. Waves in Jupiter's atmosphere observed by the Cassini ISS and CIRS instruments
- Author
-
Liming Li, F. Michael Flasar, Robert A. West, Amy A. Simon-Miller, Ulyana A. Dyudina, Carolyn C. Porco, Andrew P. Ingersoll, Richard Achterberg, Ashwin R. Vasavada, and Shawn P. Ewald
- Subjects
Atmosphere ,Jupiter ,Physics ,Solar System ,Haze ,Space and Planetary Science ,Astronomy ,Polar ,Astronomy and Astrophysics ,Zonal and meridional ,Stratosphere ,Latitude - Abstract
The Cassini Imaging Science Subsystem (ISS) and Composite Infrared Spectrometer (CIRS) reported a North Equatorial Belt (NEB) wave in Jupiter's atmosphere from optical images [Porco, C.C., and 23 colleagues, 2003. Science 299, 1541–1547] and thermal maps [Flasar, F.M., and 39 colleagues, 2004. Nature 427, 132–135], respectively. The connection between the two waves remained uncertain because the two observations were not simultaneous. Here we report on simultaneous ISS images and CIRS thermal maps that confirm that the NEB wave shown in the ISS ultraviolet (UV1) and strong methane band (MT3) images is correlated with the thermal wave in the CIRS temperature maps, with low temperatures in the CIRS maps (upwelling) corresponding to dark regions in the UV1 images (UV-absorbing particles) and bright regions in the MT3 images (high clouds and haze). The long period of the NEB wave suggests that it is a planetary (Rossby) wave. The combined observations from the ISS and CIRS are utilized to discuss the vertical and meridional propagation of the NEB wave, which offers a possible explanation for why the NEB wave is confined to specific latitudes and altitudes. Further, the ISS UV1 images reveal a circumpolar wave centered at 48.5° S (planetocentric) and probably located in the stratosphere, as suggested by the ISS and CIRS observations. The simultaneous comparison between the ISS and CIRS also implies that the large dark oval in the polar stratosphere of Jupiter discovered in the ISS UV1 images [Porco, C.C., and 23 colleagues, 2003. Science 299, 1541–1547] is the same feature as the warm regions at high northern latitudes in the CIRS 1-mbar temperature maps [Flasar, F.M., and 39 colleagues, 2004. Nature 427, 132–135]. This comparison supports a previous suggestion that the dark oval in the ISS UV1 images is linked to auroral precipitation and heating [Porco, C.C., and 23 colleagues, 2003. Science 299, 1541–1547].
- Published
- 2006
- Full Text
- View/download PDF
17. Interaction of moist convection with zonal jets on Jupiter and Saturn
- Author
-
Liming Li, Andrew P. Ingersoll, and Xianglei Huang
- Subjects
Convection ,Physics ,Length scale ,Atmospheric models ,Atmosphere of Jupiter ,Astronomy and Astrophysics ,Geophysics ,Atmospheric sciences ,Physics::Fluid Dynamics ,Atmosphere ,Jupiter ,Space and Planetary Science ,Saturn ,Barotropic fluid ,Physics::Space Physics ,High Energy Physics::Experiment ,Astrophysics::Earth and Planetary Astrophysics ,Physics::Atmospheric and Oceanic Physics - Abstract
Observations suggest that moist convection plays an important role in the large-scale dynamics of Jupiter's and Saturn's atmospheres. Here we use a reduced-gravity quasigeostrophic model, with a parameterization of moist convection that is based on observations, to study the interaction between moist convection and zonal jets on Jupiter and Saturn. Stable jets with approximately the same width and strength as observations are generated in the model. The observed zonal jets violate the barotropic stability criterion but the modeled jets do so only if the flow in the deep underlying layer is westward. The model results suggest that a length scale and a velocity scale associated with moist convection control the width and strength of the jets. The length scale and velocity scale offer a possible explanation of why the jets of Saturn are stronger and wider than those of Jupiter.
- Published
- 2006
- Full Text
- View/download PDF
18. Microwave remote sensing of Jupiter's atmosphere from an orbiting spacecraft
- Author
-
Andrew P. Ingersoll, Scott Bolton, Samuel Gulkis, L. W. Kamp, Steven Levin, Michael A. Janssen, Mark Hofstadter, and Michael Allison
- Subjects
Physics ,Solar System ,Spacecraft ,business.industry ,Astronomy ,Astronomy and Astrophysics ,Jupiter ,Atmosphere ,Depth sounding ,Space and Planetary Science ,Brightness temperature ,Radiative transfer ,Astrophysics::Solar and Stellar Astrophysics ,Astrophysics::Earth and Planetary Astrophysics ,business ,Physics::Atmospheric and Oceanic Physics ,Astrophysics::Galaxy Astrophysics ,Microwave ,Remote sensing - Abstract
Microwave remote sounding from a spacecraft flying by or in orbit around Jupiter offers new possibilities for retrieving important and presently poorly understood properties of its atmosphere. In particular, we show that precise measurements of relative brightness temperature as a function of off-nadir emission angles, combined with absolute brightness temperature measurements, can allow us to determine the global abundances of water and ammonia and study the dynamics and deep circulations of the atmosphere in the altitude range from the ammonia cloud region to depths greater than 30 bars in a manner which would not be achievable with ground-based telescopes.
- Published
- 2005
- Full Text
- View/download PDF
19. Lightning on Jupiter observed in the line by the Cassini imaging science subsystem
- Author
-
Carolyn C. Porco, Ulyana A. Dyudina, Anthony D. Del Genio, J. Barbara, Andrew P. Ingersoll, Robert A. West, and Ashwin R. Vasavada
- Subjects
Atmosphere ,Physics ,Jupiter ,Solar System ,Atmosphere of Earth ,Space and Planetary Science ,Astronomy ,Astronomy and Astrophysics ,Storm ,Lightning ,Jovian ,Line (formation) - Abstract
Night side images of Jupiter taken by the Cassini Imaging Science Subsystem (ISS) camera with the H_α filter reveal four lightning clusters; two of them are repeated observations of the same storm. All of these flashes are associated with storm clouds seen a few hours earlier on the day side of Jupiter. Some of the clouds associated with lightning do not extend to the upper troposphere. The repeated lightning observations taken 20 hr apart show that storm clouds, whose mean lifetime is ∼4 days, are electrically active during a large fraction of their lifetime. The optical power of the lightning detected with the H_α filter compared to the clear-filter power of Galileo lightning may indicate that the H_α line in the lightning spectrum is about ten times weaker than expected, consistent with a flat spectrum having no prominent H_α line. This may suggest that lightning is generated in atmospheric layers deeper than 5 bars. This, in turn, may suggest that the water abundance of the jovian interior is more than 1 × solar. Averaged over many flashes, the most powerful Cassini lightning storm emits 0.8×10^9 W in the H_α line, which implies 4×10^(10) W of broadband optical power. This is 10 times more powerful than the most intense jovian lightning observed before by Voyager 2.
- Published
- 2004
- Full Text
- View/download PDF
20. Interannual variability of Mars global dust storms: an example of self-organized criticality?
- Author
-
Alexey A. Pankine and Andrew P. Ingersoll
- Subjects
Martian ,Solar System ,Astronomy and Astrophysics ,Storm ,Mars Exploration Program ,Atmospheric sciences ,Wind speed ,Physics::Geophysics ,Space and Planetary Science ,Dust storm ,Saltation (geology) ,Physics::Space Physics ,Environmental science ,Astrophysics::Earth and Planetary Astrophysics ,Dust devil ,Astrophysics::Galaxy Astrophysics ,Physics::Atmospheric and Oceanic Physics - Abstract
Previous simulations of martian global dust storms with a simple low-order model showed the desired interannual variability of storms if one of the model parameters—the threshold wind speed for starting saltation and lifting dust from the surface—was finely tuned. In this paper we show that the fine-tuning of this parameter could be the result of negative feedback in which processes associated with global dust storms raise the threshold and small-scale processes like dust devils, which are active in years between the storms, lower the threshold.
- Published
- 2004
- Full Text
- View/download PDF
21. Monte Carlo Radiative Transfer Modeling of Lightning Observed in Galileo Images of Jupiter
- Author
-
Ulyana A. Dyudina, Shawn P. Ewald, Ashwin R. Vasavada, and Andrew P. Ingersoll
- Subjects
Physics ,Opacity ,Monte Carlo method ,Atmosphere of Jupiter ,Astronomy and Astrophysics ,Astrophysics ,Lightning ,Light scattering ,Jovian ,Jupiter ,Space and Planetary Science ,Radiative transfer ,Astrophysics::Earth and Planetary Astrophysics ,Remote sensing - Abstract
We study lightning on Jupiter and the clouds illuminated by the lightning using images taken by the Galileo orbiter. The Galileo images have a resolution of ∼25 km/pixel and are able to resolve the shape of single lightning spots, which have half widths (radii) at half the maximum intensity in the range 45–80 km. We compare the shape and width of lightning flashes in the images with simulated flashes produced by our 3D Monte Carlo light-scattering model. The model calculates Monte Carlo scattering of photons in a 3D opacity distribution. During each scattering event, light is partially absorbed. The new direction of the photon after scattering is chosen according to a Henyey–Greenstein phase function. An image from each direction is produced by accumulating photons emerging from the cloud in a small range (bins) of emission angles. The light source is modeled either as a point or a vertical line. A plane-parallel cloud layer does not always fit the data. In some cases the cloud over the light source appears to resemble cumulus clouds on Earth. Lightning is estimated to occur at least as deep as the bottom of the expected water cloud. For the six flashes studied, we find that the clouds above the lightning are optically thick (τ>5). Jovian flashes are more regular and circular than the largest terrestrial flashes observed from space. On Jupiter there is nothing equivalent to the 30–40-km horizontal flashes that are seen on Earth.
- Published
- 2002
- Full Text
- View/download PDF
22. Shear Instabilities as a Probe of Jupiter's Atmosphere
- Author
-
Andrew P. Ingersoll and Tanja Bosak
- Subjects
Physics ,Atmospheric models ,Atmosphere of Jupiter ,Galileo Probe ,Astronomy ,Astronomy and Astrophysics ,Astrophysics ,Astron ,Wavelength ,Space and Planetary Science ,Planet ,Wind shear ,Physics::Space Physics ,Astrophysics::Earth and Planetary Astrophysics ,Linear stability - Abstract
Linear wave patterns in Jupiter's clouds with wavelengths strongly clustered around 300 km are commonly observed in the planet's equatorial atmosphere (F. M. Flasar and P. J. Gierasch, 1986, J. Atmos. Sci.43, 2683–2707). We propose that the preferred wavelength is related to the thickness of an unstable shear layer within the clouds (A. P. Ingersoll and D. W. Koerner 1989, Bull. Am. Astron. Soc.21, 943). We numerically analyze the linear stability of wavelike disturbances that have nonzero horizontal phase speeds in Jupiter's atmosphere and find that, if the static stability in the shear layer is very low (but still nonnegative), a deep vertical shear layer like the one measured by the Galileo probe (D. H. Atkinson et al. 1998, J. Geophys. Res.103, 22911–22928) can generate the instabilities. The fastest growing waves grow exponentially within an hour, and their wavelengths match the observations. Close to zero values of static stability that permit the growth of instabilities are within the range of values measured by the Galileo probe in a hot spot (A. Seiff et al. 1998, J. Geophys. Res.103, 22857–22889). Our model probes Jupiter's equatorial atmosphere below the cloud deck and suggests that thick regions of wind shear and low static stability exist outside hot spots.
- Published
- 2002
- Full Text
- View/download PDF
23. Interannual Variability of Martian Global Dust Storms Simulations with a Low-Order Model of the General Circulation
- Author
-
Andrew P. Ingersoll and Alexey A. Pankine
- Subjects
Martian ,Atmospheric physics ,Meteorology ,Atmospheric circulation ,Astronomy and Astrophysics ,Storm ,Atmosphere of Mars ,Atmospheric sciences ,Space and Planetary Science ,Dust storm ,Saltation (geology) ,Physics::Space Physics ,Environmental science ,Astrophysics::Earth and Planetary Astrophysics ,Hadley cell ,Physics::Atmospheric and Oceanic Physics - Abstract
We present simulations of the interannual variability of martian global dust storms (GDSs) with a simplified low-order model (LOM) of the general circulation. The simplified model allows one to conduct computationally fast long-term simulations of the martian climate system. The LOM is constructed by Galerkin projection of a 2D (zonally averaged) general circulation model (GCM) onto a truncated set of basis functions. The resulting LOM consists of 12 coupled nonlinear ordinary differential equations describing atmospheric dynamics and dust transport within the Hadley cell. The forcing of the model is described by simplified physics based on Newtonian cooling and Rayleigh friction. The atmosphere and surface are coupled: atmospheric heating depends on the dustiness of the atmosphere, and the surface dust source depends on the strength of the atmospheric winds. Parameters of the model are tuned to fit the output of the NASA AMES GCM and the fit is generally very good. Interannual variability of GDSs is possible in the LOM, but only when stochastic forcing is added to the model. The stochastic forcing could be provided by transient weather systems or some surface process such as redistribution of the sand particles in storm generating zones on the surface. The results are sensitive to the value of the saltation threshold, which hints at a possible feedback between saltation threshold and dust storm activity. According to this hypothesis, erodable material builds up as a result of a local process, whose effect is to lower the saltation threshold until a GDS occurs. The saltation threshold adjusts its value so that dust storms are barely able to occur.
- Published
- 2002
- Full Text
- View/download PDF
24. Interpretation of NIMS and SSI Images on the Jovian Cloud Structure
- Author
-
G. E. Danielson, R. W. Carlson, Kevin H. Baines, Andrew P. Ingersoll, and Ulyana A. Dyudina
- Subjects
Physics ,Brightness ,Spectrometer ,business.industry ,Multispectral image ,Astronomy and Astrophysics ,Hot spot (veterinary medicine) ,Spectral line ,Wavelength ,Optics ,Space and Planetary Science ,Great Red Spot ,business ,Absorption (electromagnetic radiation) - Abstract
We present maps of jovian cloud properties derived from images taken simultaneously by the Galileo solid state imaging system (SSI) and the near-infrared mapping spectrometer (NIMS) at 26 visible and near-infrared wavelengths, ranging from 0.41 to 5.2μm. Three regions—the Great Red Spot (GRS), a 5-micron Hot Spot, and one of the White Ovals—were studied. We perform a principal component analysis (PCA) on the multispectral images. The principal components (PCs), also known as empirical orthogonal functions, depend only on wavelength. The first PC is that spectral function which, when multiplied by an optimally chosen number (amplitude factor) at each pixel location and subtracted from the spectrum there, minimizes the variance for the image as a whole. Succeeding PCs minimize the residual variance after the earlier PCs have been subtracted off. We find that the pixel-to-pixel variations at the different wavelengths are highly correlated, such that the first three PCs explain 91% of the variance in the spectra. Further, one can estimate the amplitudes of the first two PCs using only the four SSI wavelengths and still explain 62% of the variance of the entire spectrum. This can be an advantage when trying to classify features that are resolved in the SSI images but not in the NIMS images. The first PC in all three regions shows negative correlation between 5μm emission and reflected solar light in both atmospheric windows and the methane and ammonia absorption bands. Thus most of the bright, optically thick clouds blocking thermal emission are also extended vertically to the upper troposphere. The first PC at the GRS shows a negative correlation between the violet and all other bands except 5μm for which the correlation is positive. Thus in the GRS there is a red chromophore (absorbing in the violet, reflecting at longer wavelengths) which is associated with clouds that block 5-μm emission. There is no such correlation at the hot spot and white oval regions and therefore no chromophore associated with clouds. The second PC shows a positive correlation between the depth of the methane and ammonia absorption bands and brightness at other visible and near-IR wavelengths; there is also a negative correlation between these quantities and 5-μm emission. Thus some of the bright, optically thick clouds blocking thermal emission are deep and do not extend vertically to the upper troposphere. A color image composed using the first three PCs shows areas of unusual spectra, which appear in distinct colors. An example is the small convective stormlike cloud to the northwest of the GRS. This cloud is highly reflective at long wavelengths (4μm) and might indicate unusually large particles.
- Published
- 2001
- Full Text
- View/download PDF
25. Cold Spots in the Martian Polar Regions: Evidence of Carbon Dioxide Depletion?
- Author
-
Andrew P. Ingersoll and Benjamin P. Weiss
- Subjects
Martian ,Brightness ,Cold spot ,Astronomy and Astrophysics ,Mars Exploration Program ,Atmospheric sciences ,Condensation temperature ,chemistry.chemical_compound ,Warm front ,chemistry ,Space and Planetary Science ,Carbon dioxide ,Polar ,Environmental science ,Astrophysics::Earth and Planetary Astrophysics ,Physics::Atmospheric and Oceanic Physics - Abstract
Regions of very low, rapidly varying brightness temperatures have been observed near the martian winter poles by several spacecraft. One possibility is that the CO2 condensation temperature is lowered by depletion of CO2 in the air at the surface. We estimate the rate at which this low-molecular-weight air would disperse into the high-molecular-weight air above and show that it is generally faster than the rate of supply. This dispersal could be prevented if there is a strong temperature inversion (warm air above colder air) near the surface. Without an inversion, the entire atmospheric column could become depleted. However, depleted columns take a long time to form, and they are inconsistent with the rapid fluctuations in the cold spot locations and temperatures. Because low-altitude temperature inversions cannot be ruled out by existing observations, CO2 depletion is still a viable explanation for the martian cold spots.
- Published
- 2000
- Full Text
- View/download PDF
26. The State and Future of Mars Polar Science and Exploration
- Author
-
Bruce C. Murray, François Forget, Stephen M. Clifford, Erik W. Blake, William D. Harrison, Dorthe Dahl-Jensen, David D. Wynn-Williams, Aaron P. Zent, S. E. Wood, John F. Nye, Kenneth Lepper, James W. Rice, Daniel J. McCleese, James A. Cutts, K. E. Herkenhoff, Andrew P. Ingersoll, Fraser P. Fanale, Bruce G. Bills, Robert M. Haberle, William B. Durham, Peter C. Thomas, Benton C. Clark, Suzanne E. Smrekar, Ralph P. Harvey, David E. Smith, Jack D. Farmer, Michael H. Carr, Ellen Mosley-Thompson, R. Grard, Kumiko Gotto-Azuma, Jonathan Cameron, Philip R. Christensen, Philip B. James, David A. Paige, Stephen R. Platt, Kenneth L. Tanaka, Hugh H. Kieffer, Jeffrey S. Kargel, H. Jay Zwally, Gary D. Clow, Wendy M. Calvin, David A. Fisher, Alan D. Howard, Carol R. Stoker, J. J. Plaut, Niels Reeh, David Crisp, Jeffrey R. Barnes, Thorsteinn Thorsteinsson, Maria T. Zuber, Janus Larsen, Richard W. Zurek, and Michael C. Malin
- Subjects
Extraterrestrial Environment ,010504 meteorology & atmospheric sciences ,Climate ,Solar luminosity ,Mars ,01 natural sciences ,Astrobiology ,Atmosphere ,Impact crater ,Planet ,Dust storm ,Exobiology ,0103 physical sciences ,010303 astronomy & astrophysics ,0105 earth and related environmental sciences ,Martian ,Ice ,Astronomy and Astrophysics ,Mars Exploration Program ,Atmosphere of Mars ,Carbon Dioxide ,Space Flight ,Cold Climate ,13. Climate action ,Space and Planetary Science ,Geology - Abstract
As the planet's principal cold traps, the martian polar regions have accumulated extensive mantles of ice and dust that cover individual areas of approximately 10(6) km2 and total as much as 3-4 km thick. From the scarcity of superposed craters on their surface, these layered deposits are thought to be comparatively young--preserving a record of the seasonal and climatic cycling of atmospheric CO2, H2O, and dust over the past approximately 10(5)-10(8) years. For this reason, the martian polar deposits may serve as a Rosetta Stone for understanding the geologic and climatic history of the planet--documenting variations in insolation (due to quasiperiodic oscillations in the planet's obliquity and orbital elements), volatile mass balance, atmospheric composition, dust storm activity, volcanic eruptions, large impacts, catastrophic floods, solar luminosity, supernovae, and perhaps even a record of microbial life. Beyond their scientific value, the polar regions may soon prove important for another reason--providing a valuable and accessible reservoir of water to support the long-term human exploration of Mars. In this paper we assess the current state of Mars polar research, identify the key questions that motivate the exploration of the polar regions, discuss the extent to which current missions will address these questions, and speculate about what additional capabilities and investigations may be required to address the issues that remain outstanding.
- Published
- 2000
- Full Text
- View/download PDF
27. The 1997 Spring Regression of the Martian South Polar Cap: Mars Orbiter Camera Observations
- Author
-
Andrew P. Ingersoll, Peter C. Thomas, Kenneth S. Edgett, G. E. Danielson, Philip B. James, Michael C. Malin, J. Veverka, Alfred S. McEwen, B. A. Cantor, William K. Hartmann, Merton E. Davies, L. A. Soderblom, and Michael H. Carr
- Subjects
Martian ,Water on Mars ,Astronomy and Astrophysics ,Evidence of water on Mars from Mars Odyssey ,Mars Exploration Program ,Exploration of Mars ,Geodesy ,law.invention ,Astrobiology ,Orbiter ,Space and Planetary Science ,law ,Mars Orbiter Laser Altimeter ,Longitude ,Geology - Abstract
The Mars Orbiter cameras (MOC) on Mars Global Surveyor observed the south polar cap of Mars during its spring recession in 1997. The images acquired by the wide angle cameras reveal a pattern of recession that is qualitatively similar to that observed by Viking in 1977 but that does differ in at least two respects. The 1977 recession in the 0° to 120° longitude sector was accelerated relative to the 1997 observations after L_S = 240°; the Mountains of Mitchel also detached from the main cap earlier in 1997. Comparison of the MOC images with Mars Orbiter Laser Altimeter data shows that the Mountains of Mitchel feature is controlled by local topography. Relatively dark, low albedo regions well within the boundaries of the seasonal cap were observed to have red-to-violet ratios that characterize them as frost units rather than unfrosted or partially frosted ground; this suggests the possibility of regions covered by CO_2 frost having different grain sizes.
- Published
- 2000
- Full Text
- View/download PDF
28. Ejecta Pattern of the Impact of Comet Shoemaker–Levy 9
- Author
-
Alexey A. Pankine and Andrew P. Ingersoll
- Subjects
Physics ,Mass distribution ,Comet ,Astronomy ,Astronomy and Astrophysics ,Lateral expansion ,Rotation ,Jovian ,Plume ,Jupiter ,Space and Planetary Science ,Astrophysics::Solar and Stellar Astrophysics ,Astrophysics::Earth and Planetary Astrophysics ,Ejecta - Abstract
The collision of Comet Shoemaker–Levy 9 (SL 9) with Jupiter created crescent-shaped ejecta patterns around impact sites. Although the observed impact plumes rose through a similar height of ∼3000 km, the radii of the created ejecta patterns differ from impact to impact and generally are larger for larger impacts. The azimuthal angle of the symmetry axis of the ejecta pattern is larger than that predicted by the models of oblique impacts, due to the action of the Coriolis force that rotates ejecta patterns counterclockwise from the south. We study the formation of ejecta patterns using a simple model of ballistic plume above a rotating plane. The ejected particles follow ballistic trajectories and slide horizontally for about an hour after reentry into the jovian atmosphere. The lateral expansion of the plume is stopped by the friction force, which is assumed to be proportional to the square of the horizontal velocity. Two different mass–velocity distributions used in the simulations produce qualitatively similar results. The simulated ejecta patterns fit very well the “crescents” observed at the impact sites. The sizes and azimuthal angles of symmetry axis of ejecta patterns depend on a parameterL, which has dimension of length and is related to the mass of the fragment. Thus more massive impacts produce larger ejecta patterns that are rotated through a wider angle.
- Published
- 1999
- Full Text
- View/download PDF
29. Imaging Jupiter's Aurora at Visible Wavelengths
- Author
-
Claudia Alexander, C. Anger, W. Kent Tobiska, Kenneth P. Klaasen, Blane Little, Scott Bolton, Ashwin R. Vasavada, and Andrew P. Ingersoll
- Subjects
Physics ,Brightness ,Flux tube ,business.industry ,Equator ,Astronomy and Astrophysics ,Astrophysics ,Jovian ,Optics ,Space and Planetary Science ,Planet ,Radiance ,Longitude ,business ,Image resolution - Abstract
On November 9, 1996 and again on April 2, 1997, the Galileo spacecraft's Solid State Imaging (SSI) camera targeted the northern auroral region of Jupiter. These observations represent (i) the first spatially resolved images of the jovian auroral oval either at visible wavelengths or on the nightside of the planet, (ii) the first image at visible wavelengths of an auroral footprint of the Io Flux Tube (IFT), (iii) the first unambiguous detection at visible wavelengths of auroral emission on the jovian limb, and (iv) the first images of the aurora with spatial resolution below 100 km per pixel (46 and 35 km, respectively). Relative to many prior expectations, the visible aurora is (i) lower in altitude, (ii) associated with magnetic field lines that cross the equator closer to the planet, and (iii) more variable in time and space. The 1996 images used a clear (broadband) filter, while the 1997 images used both the clear filter and five narrower filters over wavelengths ranging from violet to 968 nm. The filtered images imply that the visible auroral emission contains atomic hydrogen lines, although there is also a continuum component. We were able to position the aurora in three-dimensional space and found the limb emission to be ∼240 km above the surface of a standard (P≈ 1 bar) reference ellipsoid. Our most accurate analysis of the equatormost part of the oval placed it at 54.5° planetocentric latitude and 168° west longitude. Combined with the latest magnetic field models, our results imply that the particles that cause the aurora originate in Jupiter's equatorial plane ∼13 R_J from the center of the planet. The oval was brighter and wider in the 1996 images than in the 1997 images. The broadband radiance of a typical place on the oval as seen directly overhead varied from ∼80 kR in 1997 to ∼300 kR in 1996. Our estimates of the full width of the oval varied from under 500 km to over 8000 km, partly depending on the signal-to-noise ratio of the image. The radiated power per unit length along the oval ranged from ∼60 to ∼700 W/m, with the associated radiated power from the entire oval varying from ∼109 to ∼9 × 10^(10) W. Appreciable auroral emission also occurred both north and south of the main oval. One image contains the northern footprint of the IFT, which appears as a central ellipse with a tail of emission that lies downstream with respect to the plasma flow past Io. The central ellipse is ∼1200 km downstream by ∼500 km cross stream. The IFT is comparable in brightness to the nearby auroral oval (∼250 kR) and has a total radiated power of ∼3 × 10^8 W.
- Published
- 1998
- Full Text
- View/download PDF
30. Galileo Imaging of Jupiter's Atmosphere: The Great Red Spot, Equatorial Region, and White Ovals
- Author
-
Todd J. Jones, Andrew P. Ingersoll, H. Herbert Breneman, Ashwin R. Vasavada, David A. Senske, James M. Kaufman, Kenneth P. Klaasen, E. DeJong, K. Magee, Maureen Bell, Peter J. Gierasch, Don Banfield, Glenn S. Orton, and Michael J. S. Belton
- Subjects
Jupiter ,Space and Planetary Science ,Anticyclone ,Cloud height ,Galileo Probe ,Great Red Spot ,Astronomy ,Astronomy and Astrophysics ,Hot spot (veterinary medicine) ,Context (language use) ,Great Dark Spot ,Geology - Abstract
During the first six orbits of the Galileo spacecraft's prime mission, the Solid State Imaging (SSI) system acquired multispectral image mosaics of Jupiter's Great Red Spot, an equatorial belt/zone boundary, a “5-μm hot spot” similar to the Galileo Probe entry site, and two of the classic White Ovals. We present mosaics of each region, approximating their appearance at visible wavelengths and showing cloud height and opacity variations. The local wind field is derived by tracking cloud motions between multiple observations of each region with time separations of roughly 1 and 10 hr. Vertical cloud structure is derived in a companion paper by Banfieldet al. (Icarus135, 230–250). Galileo's brief, high-resolution observations complement Earth-based and Voyager studies and offer local meteorological context for the Galileo Probe results. Our results show that the dynamics of the zonal jets and large vortices have changed little since Voyager, with a few exceptions. We detect a cyclonic current within the center of the predominantly anticyclonic Great Red Spot. The zonal velocity difference between 0° S and 6° S has increased by 20 m sec^(−1). We measure a strong northeast flow approaching the hot spot. This flow indicates either massive horizontal convergence or the presence of a large anticyclonic vortex southeast of the hot spot. The current compact arrangement of two White Ovals and a cyclonic structure greatly perturbs the zonal jets in that region.
- Published
- 1998
- Full Text
- View/download PDF
31. Jupiter's Cloud Structure from Galileo Imaging Data
- Author
-
Eugene A. Ustinov, Don Banfield, M. J. S. Belton, Robert A. West, Peter J. Gierasch, Ashwin R. Vasavada, M. Bell, and Andrew P. Ingersoll
- Subjects
Haze ,Scattering ,Astronomy and Astrophysics ,Scale height ,Astrophysics ,Atmospheric sciences ,Aerosol ,Troposphere ,Wavelength ,Space and Planetary Science ,Great Red Spot ,Astrophysics::Earth and Planetary Astrophysics ,Optical depth ,Geology - Abstract
The vertical structure of aerosols on Jupiter is inferred from data obtained by the NASA Galileo Solid State Imaging system during the first six orbits of the spacecraft. Images at 889 nm (a strong methane band), 727 nm (a weaker methane band), and 756 nm (continuum) taken at a variety of lighting and viewing angles are used. The images are displayed and described in the companion paper by Vasavadaet al.(1998,Icarus135, 265–275). Conservative scattering cloud particles with laterally uniform single scattering properties are assumed in the analysis and are shown to be consistent with the data at these wavelengths. Particles are bright, and the darkest locations on Jupiter correspond to the smallest optical thickness of aerosols. Optical depths and vertical positions of aerosol layers vary from place to place and are the retrieved quantities in the analysis. Only mid and low latitudes are sampled in this data set. A stratospheric haze with an optical depth of roughly a tenth and an upper tropospheric haze with an optical depth of 2 to 6 exist over all regions. Both are consistent with previous conclusions based on data of lower spatial resolution (e.g., Westet al.1986,Icarus65, 161–217). The new data show that these layers contain little lateral structure on scales smaller than the planetary jets. On scales of the jets and ovals, the top and bottom of the upper tropospheric haze vary in elevation. The concentration of particles (optical depth per pressure interval) varies less than does the total optical depth. Near the base of the upper tropospheric haze is a third cloud component, usually at pressure p= 750 ± 200 mb, which is less than a scale height in geometric thickness. Its optical depth varies from zero to about 20 on regional scales and often varies by 50% on scales of a few tens of kilometers. Optical depth variations in this cloud are the principal cause of the features in Jupiter's atmosphere seen at red and longer wavelengths. It is probably composed of ammonia. The expected NH_4SH cloud has not been identified in this work, perhaps because it exists only at locations where it is concealed beneath higher clouds. Our retrievals also cannot rule out a pervasive deep haze without small-scale structure. Finally, in one region northwest of the Great Red Spot, a deeper cloud is identified. Parts of it lie at a pressure greater than four bars. It is associated with a rapidly changing storm system with optical depth of several tens (or more) and a range of cloud heights between p> 4 bars top∼ 400 mb. It is probably composed of water.
- Published
- 1998
- Full Text
- View/download PDF
32. Interpretation of Galileo Probe Data and Implications for Jupiter's Dry Downdrafts
- Author
-
Adam P. Showman and Andrew P. Ingersoll
- Subjects
Atmosphere ,Jupiter ,Convection ,Meteorology ,Space and Planetary Science ,Downwelling ,Potential density ,Galileo Probe ,Astronomy and Astrophysics ,Hot spot (veterinary medicine) ,Atmospheric sciences ,Geostrophic wind ,Geology - Abstract
The Galileo probe found the jovian abundance of H_2S to be 30% solar at the 8 bar level, while the abundance of water was less than 3% solar at 12 bars. From 8 to 20 bars, H_2S increased to three times solar, and water apparently increased as well. Since H_2S and water condense at 2 and 5 bars, respectively, the probe probably entered a dry downdraft, wherein dry air above 2 bars is advected to 12 bars or deeper (Owen et al. 1996,Eos (Spring Suppl.) 77, S171). This is consistent with the fact that the probe entered the south edge of a 5-μm hot spot, a local region of Jupiter's atmosphere known from spectral modeling to be unusually low in cloud abundance (Orton et al. 1996, Science 272, 839). We use basic physical constraints to address three problems raised by Galileo probe data. First, it is unclear how the hypothesized downdraft remains dry, since simple models of convection preclude dessication below the 2- and 5-bar condensation levels. We suggest that to suppress moist plumes from below, the downdraft must be of low density below 5 bars and hence thermally indirect, requiring mechanical forcing from other parts of the atmosphere. Second, if geostrophic balance holds, the Galileo probe winds imply that the hot spot (north of the probe site) contains a stable layer from 1 to 5 bars; this is inconsistent with a downwelling, since downwellings should be adiabatic below 2 bars due to the low radiative flux divergence. We show that when the centripetal acceleration of curving parcel trajectories is included in the force balance, however, a variety of density profiles is possible within the hot spot (depending on the radius of curvature of the winds). The most plausible profile implies that the hot spot is nearly dry adiabatic and that the equatorial zone south of the probe site is stable from 2 to 6 bars, suggesting moist adiabatic upwellings with a water abundance of 1–2 times solar. This is consistent with Galileo and Voyager images suggesting upwelling at the equator. The profile further implies that from 1 to 5 bars the hot spot is denser than the equatorial zone south of the probe site. Third, probe data indicate that NH_3 increased with depth below 1 bar and became constant by 8 bars, H_2S began increasing below 8 bars and leveled off by 16 bars, while water only began increasing below 12 bars and was still increasing with depth at 20 bars. We propose that lateral mixing along isopycnals (surfaces of constant potential density) could produce the observed pattern; alternatively, the downwelling might consist of column stretching, so that the NH_3, NH_4SH, and water lifting condensation levels were pushed to 8, 16, and >20 bars, respectively. In either case, the simplest form of this model requires the downdraft to be less dense than the surroundings from 0.5 to 20 bars. In its simplest form, this model is therefore incompatible with our favored interpretation of the winds; more detailed studies will be necessary to resolve the problem.
- Published
- 1998
- Full Text
- View/download PDF
33. Martian Weather Correlation Length Scales
- Author
-
Anthony D. Toigo, Andrew P. Ingersoll, Don Banfield, and David A. Paige
- Subjects
Martian ,Data assimilation ,Space and Planetary Science ,Brightness temperature ,Rossby radius of deformation ,Environmental science ,Astronomy and Astrophysics ,Mars Exploration Program ,Atmospheric temperature ,Atmospheric sciences ,Longitude ,Latitude - Abstract
Spring and fall equinox Viking infrared thermal mapper 15-μm channel atmospheric brightness temperature (T15) observations are used to estimate the weather correlation length scale of Mars in the pressure range 0.5–1 mbar. The results provide a better understanding of martian atmospheric dynamics, a benchmark for validating martian general circulation models (GCMs), a guide to the optimal placement of a network of landers, and information for use in data assimilation efforts for orbiters and landers. Observations of atmospheric temperature are used to compute an atmospheric mean state as a function of time-of-day, latitude, longitude, and altitude, which is then subtracted from the observations to yield weather temperature residuals. These residuals are correlated with each other to determine (1) the weather temperature correlation length scale (∼1000 km) as a function of latitude and (2) the weather temperature variance (∼4 K^2 global average for L_S∼ 0°, ∼3 K^2 for L_S ∼ 180°). Good general agreement is found in comparing the length scales to the Rossby radius of deformation and to inferences made from other data sets. The weather temperature variance results are also compared with GCM results, yielding satisfactory agreement, with some differences in the magnitudes of the variances.
- Published
- 1996
- Full Text
- View/download PDF
34. Baroclinic Instability in the Interiors of the Giant Planets: A Cooling History of Uranus?
- Author
-
Andrew P. Ingersoll and Richard Holme
- Subjects
Physics ,Gas giant ,Baroclinity ,Equator ,Uranus ,Astronomy ,Astronomy and Astrophysics ,Astrophysics ,Instability ,Space and Planetary Science ,Neptune ,Planet ,Astrophysics::Earth and Planetary Astrophysics ,Internal heating - Abstract
We propose a quasigeostrophic, baroclinic model for heat transport within the interior of a stably stratified Jovian planet, based on motion in thin cylindrical annuli. Density decreases from the center outward and is zero at the surface of the planet. In the homogeneous case (no core), we find instability for the poles hotter than the equator, but not for the reverse. If the motion is bounded by an impenetrable core, instability occurs for both cases. Much of the behavior can be explained by analogy to conventional baroclinic instability theory. Motivated by our results, we explore a possible connection between the highly inclined rotation axis of Uranus and its anomalously low surface heat flux. We assume that the planets formed hot. Our conjecture is that heat was efficiently convected outwards by baroclinic instability in Uranus (with the poles hotter than the equator), but not in the other three Jovian planets. The surface temperature was higher for the stably stratified case (Uranus), leading to a higher rate of infrared emission and faster cooling. Therefore, we propose that Uranus lost its internal heat sooner than Neptune because baroclinic motions, permitted by its inclination to the sun, were able to extract its internal heat while the surface was still warm.
- Published
- 1994
- Full Text
- View/download PDF
35. Hubble space telescope observations of the 1990 equatorial disturbance on Saturn: Images, albedos, and limb darkening
- Author
-
Eric M. De Jong, James A. Westphal, John Caldwell, William A. Baum, Christopher D. Barnet, G. Edward Danielson, and Andrew P. Ingersoll
- Subjects
Disturbance (geology) ,Space and Planetary Science ,Limb darkening ,Planet ,Saturn ,Astronomy ,Astronomy and Astrophysics ,Meridian (astronomy) ,Albedo ,Longitude ,Geology ,Latitude - Abstract
In September 1990 a major equatorial eruption on Saturn produced a disturbance that spread in longitude until it completely girdled the planet. We report here on 150 images recorded in six passbands with the Wide Field/Planetary Camera (WF/PC) aboard the Hubble Space Telescope (HST) on 17 and 18 November 1990. For comparison, we used HST-WF/PC observations of Saturn obtained in three colors on 26 August 1990 before the onset of the disturbance, and in six colors on 5 and 6 June 1991 when almost no evidence of the disturbance remained. At both of those times, the equatorial belt was “normal” in appearance. Four of the passbands (with mean wavelengths of 336, 435, 546, and 716 nm) were selected for photometric analysis, and a patch of the B ring near the central meridian was used for photometric calibration. Using deconvolved images from all three epochs of observation, we measured reflectivities (I/F) of the disk along parallels of latitude as a function of longitudinal distance from the central meridian and also along the central meridian as a function of latitude from 0° to 90°. The longitudinal measurements cover essentially the whole visible disk; they were made at 1° intervals of planetographic latitude from 0° to 80°, and the results are expressed in terms of Minnaert coefficients k and Minnaert albedos (I/F)_0. We find that the cloud particles associated with the disturbance must differ in character from those that normally make up the visible cloud deck on Saturn. They were brighter and bluer, they had greater limb darkening, and their limb darkening was spectrally more neutral. The mutual relationship of those properties is such that features which stand out strongly near the meridian fade to invisibility when near the limb.
- Published
- 1992
- Full Text
- View/download PDF
36. Stability of polar frosts in spherical bowl-shaped craters on the Moon, Mercury, and Mars
- Author
-
Andrew P. Ingersoll, Bruce C. Murray, and Tomas Svitek
- Subjects
Solar constant ,Lunar craters ,Radiative equilibrium ,Impact crater ,Geology of the Moon ,Heat flux ,Space and Planetary Science ,Emissivity ,Astronomy and Astrophysics ,Mars Exploration Program ,Atmospheric sciences ,Geology ,Astrobiology - Abstract
Following Svitek (Martian Water Frost: Control of Global Distribution by Small-Scale Processes, Ph.D Thesis, California Institute of Technology, 1992), analytic solutions are presented for the effective albedo, the effective emissivity, and the radiative equilibrium temperature in the shadowed portions of a spherical bowl-shaped crater. The model assumes that the surface is a Lambert scatterer with visual albedo and infrared emissivity each independent of wavelength across their respective spectral ranges. Absorption, emission, and multiple scattering from the walls of the crater are treated rigorously to all orders. For airless bodies whose surfaces are in radiative equilibrium, all shadowed portions of any individual crater have the same temperature, whose value depends on four quantities: the insolation (product of the solar constant and the sine of the solar elevation angle), the depth/diameter ratio of the crater, the visual albedo, and the infrared emissivity. As long as the crater is deep enough to have shadows, the lowest temperatures are for the shallowest crater—those with the smallest depth/diameter ratio. The model is applied first to the Moon and Mercury using a depth/diameter ratio of 0.2, which is typical of the lunar highlands according to Pike (Geophys. Res. Lett. 1, 291–294 (1974); in Impact and Explosion Cratering (Roddy et al., Eds.), pp. 489–509, Pergamon, New York, 1977). For Mercury and the Moon, temperatures in shadows in polar craters are below 102 K, so the sublimation rate of water ice calculated according to the model of Watson et al. (J. Geophys. Res. 66, 3033-3015 (1961)) is less than 1 cm per byr. The latitudinal extent of the cold zone on the Moon is greater than that on Mercury, although temperatures at the poles of the two planets are similar. The other application is to polar frosts on Mars. Illuminated water frosts in radiative equilibrium grow rougher, because the average temperature of a depression is greater than that of flat ground. Subliming CO_2 frosts, which are always at the same temperature, grow rougher at low solar elevation angles because the heat flux absorbed by a depression is greater than that for a flat surface. At high insolation rates (high Sun near perihelion) the average heat flux to a depression is less than for a flat surface. The latter evaporates faster, which makes the average surface smoother and leads to a high average albedo. This behavior helps explain the fact that the south CO_2 cap, which receives its greatest insolation near perihelion, has a higher effective albedo and therefore can survive the summer, whereas the north CO_2 cap has a lower effective albedo and disappears each year around summer solstice.
- Published
- 1992
- Full Text
- View/download PDF
37. Book review
- Author
-
Andrew P. Ingersoll
- Subjects
Atmospheric physics ,Space and Planetary Science ,Philosophy ,Art history ,Astronomy and Astrophysics - Published
- 1992
- Full Text
- View/download PDF
38. Seasonal buffering of atmospheric pressure on Mars
- Author
-
Daniel Dzurisin and Andrew P. Ingersoll
- Subjects
ICARUS ,Martian ,Atmospheric models ,Atmospheric pressure ,Astronomy and Astrophysics ,Mars Exploration Program ,Atmosphere of Mars ,Atmospheric sciences ,Regolith ,Atmosphere ,Space and Planetary Science ,Physics::Space Physics ,Environmental science ,Martian polar ice caps ,Astrophysics::Earth and Planetary Astrophysics ,Physics::Atmospheric and Oceanic Physics - Abstract
An isothermal reservoir of carbon dioxide in gaseous contact with the Martian atmosphere would reduce the amplitude and advance the phase of global atmospheric pressure fluctuations caused by seasonal growth and decline of polar CO2 frost caps. Adsorbed carbon dioxide in the upper roughly 10 m of Martian regolith is sufficient to buffer the present atmosphere on a seasonal basis. Available observations and related polar cap models do not confirm or refute the operation of such a mechanism. Implications for the amplitude and phase of seasonal pressure fluctuations are subject to direct test by the upcoming Viking mission to Mars.
- Published
- 1975
- Full Text
- View/download PDF
39. Lateral inhomogeneities in the Venus atmosphere: Analysis of thermal infrared maps
- Author
-
Andrew P. Ingersoll and Glenn S. Orton
- Subjects
biology ,Terminator (solar) ,Antisolar point ,Equator ,Subsolar point ,Astronomy and Astrophysics ,Venus ,Atmospheric sciences ,biology.organism_classification ,Geodesy ,Latitude ,Atmosphere of Venus ,Space and Planetary Science ,Physics::Space Physics ,Astrophysics::Solar and Stellar Astrophysics ,Astrophysics::Earth and Planetary Astrophysics ,Physics::Atmospheric and Oceanic Physics ,Geology ,Intensity (heat transfer) - Abstract
The thermal infrared maps of Venus published by Murray, Wildey, and Westphal (1963) and Westphal, Wildey, and Murray (1965) have been analyzed systematically in order to separate the observed intensity into a limb-darkening component and a solar-associated component representing fixed patterns of intensity corotating with the earth and sun, respectively. Interesting new results are obtained for the solar-associated component. Regions near the subsolar point and the poles are not covered in the original maps or in the analysis. The solar-associated pattern of intensity is very nearly symmetric about the equator. In both northern and southern hemispheres, an intensity minimum seems to occur near the morning terminator at middle to high latitudes, slightly beyond the limit of the maps. An intensity maximum occurs on the equator slightly to the east of the antisolar point. Three broad ridges of relatively high intensity radiate away from this point, one pointing to the west along the equator, the others pointing to the northeast and southeast, respectively. The eastward tilt of the latter two ridges may indicate that horizontal exchange is important in maintaining the equatorial maximum of zonal momentum which is associated with the 4-day circulation of the Venus atmosphere.
- Published
- 1974
- Full Text
- View/download PDF
40. Solar heating and internal heat flow on Jupiter
- Author
-
Carolyn Porco and Andrew P. Ingersoll
- Subjects
Convection ,Materials science ,Convective heat transfer ,Astronomy and Astrophysics ,Mechanics ,Heat transfer coefficient ,Atmospheric sciences ,Heat flux ,Space and Planetary Science ,Latent heat ,Physics::Space Physics ,Heat transfer ,Astrophysics::Solar and Stellar Astrophysics ,Astrophysics::Earth and Planetary Astrophysics ,Thermosiphon ,Internal heating ,Physics::Atmospheric and Oceanic Physics - Abstract
Models of convection in Jupiter's interior are studied to test the hypothesis that internal heat balances the absorbed sunlight at each latitude. Such a balance requires that the ratio of total emitted heat to absorbed sunlight be above a critical value 4/π ≈ 1.27. The necessary horizontal heat transport then takes place in the interior instead of in the atmosphere. Regions of stable stratification can arise in the interior owing to the effects of solar heating and rotation. In such regions, upward heat transfer takes place on sloping surfaces, as in the Earth's atmosphere, provided there are horizontal temperature gradients. Potential temperature gradients are found to be small, and the time constant for the pattern to reach equilibrium is found to be short compared to the age of the solar system. It is suggested that Jupiter and Saturn owe their axisymmetric appearance to internal heat flow, which eliminates differential heating in the atmosphere that would otherwise drive meridional motions.
- Published
- 1978
- Full Text
- View/download PDF
41. Supersonic meteorology of Io: Sublimation-driven flow of SO2
- Author
-
Andrew P. Ingersoll, Michael Summers, and Steve G. Schlipf
- Subjects
Materials science ,Atmospheric pressure ,Meteorology ,Vapor pressure ,Turbulence ,Subsolar point ,Astronomy and Astrophysics ,Atmospheric temperature ,Atmospheric sciences ,law.invention ,Boundary layer ,Space and Planetary Science ,law ,Hydrostatic equilibrium ,Choked flow - Abstract
The horizontal flow of SO_2 gas from day side to night side of Io is calculated. The surface is assumed to be covered by a frost whose vapor pressure at the subsolar point is orders of magnitude larger than that on the night side. Temperature of the frost is controlled by radiation. The flow is hydrostatic and turbulent, with velocity and entropy per particle independent of height. The vertically integrated conservation equations for mass, momentum, and energy are solved for atmospheric pressure, temperature, and horizontal velocity as functions of solar zenith angle. Formulas from boundary layer theory govern the interaction between atmosphere and surface. The flow becomes supersonic as it expands away from the subsolar point, as in the theory of rocket nozzles and the solar wind. Within 35° of the subsolar point atmospheric pressure is less than the frost vapor pressure, and the frost sublimes. Elsewhere, atmospheric pressure is greater than the frost vapor pressure, and the frost condenses. The two pressures seldom differ by more than a factor of 2. The sublimation rate at the subsolar point is proportional to the frost vapor pressure, which is a sensitive function of temperature. For a subsolar temperature of 130°K, the sublimation rate is 10^(15) molecules/cm^2/sec. Diurnally averaged sublimation rates at the equator are comparable to the 0.1 cm/year resurfacing rate required for burial of impact craters. At the poles where both the vapor pressures and atmospheric pressures are low, the condensation rates are 100 times smaller. Surface pressures near the terminator are generally too low to account for the ionosphere discovered by Pioneer 10. The possibility of a noncondensable gas in addition to SO_2 must be seriously considered.
- Published
- 1985
- Full Text
- View/download PDF
42. Io meteorology: How atmospheric pressure is controlled locally by volcanos and surface frosts
- Author
-
Andrew P. Ingersoll
- Subjects
Atmosphere ,Atmospheric pressure ,Meteorology ,Space and Planetary Science ,Vapor pressure ,Subsolar point ,Astronomy and Astrophysics ,Frost (temperature) ,Scale height ,Atmospheric sciences ,Geology ,Latitude ,Plume - Abstract
The hydrodynamic model of A. P. Ingersoll, M. E. Summers, S. G. Schilpf, 1985, Icarus 64, 375–390 is modified to include the effects of nonuniform surface properties as revealed in recent observational studies. The observations indicate that SO_2 frost is concentrated in a band within 30° latitude of the equator covering 270° of longitude, and that the darker surface at midlatitude is warmer than the frost at times of maximum insolation. The approach to the hydrodynamics has been to make the model simpler and more versatile. Solutions are now obtained in closed analytic form. We calculate atmospheric pressures, horizontal winds, sublimation rates, and condensation rates for a wide variety of surface conditions-patchy and continuous frost cover, volcanic venting (treated as a source of mass distributed over the plume area), discontinuities of surface temperature, subsurface cold trapping, and insolation propagation into the frost. The principal new concept is the horizontal averaging length L, which is equal to √2πH/α where H is the atmospheric scale height and α is the sticking coefficient. If frost is present, either on the surface or just below it, then each area of dimension L (of order 100 km) determines its own atmospheric pressure. Away from the volcanic plumes, the pressure follows the local vapor pressure. Inside the plumes the pressure is higher, e.g., a volcanic source of 10^6 kg/sec spread over an area 300 km in radius raises the local pressure by an amount equal to the vapor pressure of a frost at 124 K. The key unknowns for Io are the strength of the volcanic sources and the temperature of frost near the subsolar point. If the frost is cold (below 110 K) then the only substantial atmospheres are beneath the plumes. If the frost is warm (e.g. 125 K) then there is also a substantial atmosphere near the subsolar point. The equatorial frost deposits are then losing net mass at a substantial rate (e.g., 0.1 cm/year for a subsolar frost temperature of 125 K).
- Published
- 1989
- Full Text
- View/download PDF
43. Atmospheric tides and the rotation of Venus I. Tidal theory and the balance of torques
- Author
-
Anthony R. Dobrovolskis and Andrew P. Ingersoll
- Subjects
Physics ,biology ,Atmospheric tide ,Retrograde motion ,Astronomy and Astrophysics ,Venus ,Crust ,Atmospheric sciences ,biology.organism_classification ,Atmosphere of Venus ,Gravitational field ,Space and Planetary Science ,Planet ,Tidal force - Abstract
Insolation absorbed by the surface of Venus is quickly redeposited at the bottom of the atmosphere. This periodic heating causes mass to flow away from the warm afternoon side of the planet and into the cooler morning region. The Sun's gravitational field exerts a torque on this atmospheric tide tending to accelerate the retrograde zonal circulation. When this torque is transmitted to the crust, it can balance the despinning effect of tides in the body of Venus. The slow retrograde rotation of Venus may be a steady state among tides in the atmosphere, tides in the solid body, and possibly the influence of the Earth.
- Published
- 1980
- Full Text
- View/download PDF
44. Motions in the Interiors and atmospheres of Jupiter and Saturn
- Author
-
Ron L. Miller and Andrew P. Ingersoll
- Subjects
Physics ,Astronomy and Astrophysics ,Mechanics ,Atmospheric sciences ,Jupiter ,Amplitude ,Space and Planetary Science ,Planet ,Normal mode ,Barotropic fluid ,Saturn ,Astrophysics::Earth and Planetary Astrophysics ,Phase velocity ,Adiabatic process ,Physics::Atmospheric and Oceanic Physics - Abstract
The low-frequency motions in a rotating, adiabatic, inviscid fluid planet are barotropic, quasigeostrophic, and quasi-columnar. The only steady motions are differentially rotating cylinders in which zonal velocity ū is a function of cylindrical radius r. Projected onto the planetary surface the limiting curvature at which the flow becomes unstable is negative; its amplitude is three to four times the amplitude for thin atmospheres, for planets in which density decreases linearly to zero at the surface. This result, derived first by A.P. Ingersoll and D. Pollard (1982, Icarus 52, 62–80) for low zonal wavenumber perturbations, is shown to hold for all quasi-columnar perturbations. When ū = 0 the small amplitude motions are oscillatory. The lowest mode, as regards structure parallel to the axis, propagates eastward with a speed proportional to (wavelength)^2. Both the barotropic stability criterion and the phase speed of the normal mode oscillations have features in common with Jupiter and Saturn observations, although the test is inconclusive with current data and theory.
- Published
- 1986
- Full Text
- View/download PDF
45. Images of Jupiter from the pioneer 10 and pioneer 11 infrared radiometers: A comparison with visible and 5-μm images
- Author
-
Richard J. Terrile, Glenn S. Orton, Stephen R. Walton, and Andrew P. Ingersoll
- Subjects
Jupiter ,Brightness ,Altitude ,Space and Planetary Science ,Infrared ,Planet ,Cloud cover ,Astronomy ,Astronomy and Astrophysics ,Albedo ,Geology ,Latitude - Abstract
All of the data acquired at Jupiter by the Infrared Radiometers on board Pioneers 10 and 11 are presented in the form of images with geometric control. The images are compared with 5-μm and visible images taken in the same time frame. The association of dark (blue or brown) and light (white or red) areas with warm and cool areas (at 5, 20, and 45 μm) respectively, extends to nearly all features observed on the planet. Where the normal association of light and dark visible markings with the zonal velocity breaks down (e.g., at the latitude of the South Equatorial Belt during the Pioneer encounters), the infrared emission seems to follow the visible cloud structure rather than the zonal velocity structure. Exceptions to the general rule involve 20-μm radiation, which reflects conditions in the altitude range 0.1–0.3 bar. For example, a comparison between Pioneer 10 and 11 images suggests that the South Equatorial Belt became brighter at 20 μm, but remained constant at other wavelengths between the two encounters.
- Published
- 1981
- Full Text
- View/download PDF
46. Pioneer 10 and 11 observations and the dynamics of Jupiter's atmosphere
- Author
-
Andrew P. Ingersoll
- Subjects
Convection ,Physics ,Baroclinity ,Equator ,Atmosphere of Jupiter ,Northern Hemisphere ,Astronomy and Astrophysics ,Atmospheric sciences ,Jupiter ,Heat flux ,Space and Planetary Science ,Barotropic fluid ,Physics::Space Physics ,Astrophysics::Earth and Planetary Astrophysics ,Physics::Atmospheric and Oceanic Physics - Abstract
Three new results of the Pioneer 10 and 11 mission are discussed. The first is that effective temperature is the same at the poles and equator in spite of the large differences in solar energy deposition. This is consistent with theories of convection which suggest that an extremely small equator-to-pole temperature difference at the level of infrared emission could suppress the internal heat flux at the equator relative to the pole by an amount sufficient to balance the difference in solar energy deposition. The second result is that the effective temperature of belts is 3 to 4K greater than that of zones, which is almost exactly accounted for by the lower albedo of belts. This result cannot be interpreted uniquely, but is consistent with a model in which the internal heat flux is the same under belts and zones, and the horizontal atmospheric heat flux is zero. The third observation provides evidence of instability along the south edges of zones in the northern hemisphere. These are the latitudes of minimum prograde velocity, where instability is most likely to occur in a barotropic fluid, as pointed out by Ingersoll and Cuzzi (1969) . A more realistic baroclinic stability analysis suggests instability at these same latitudes.
- Published
- 1976
- Full Text
- View/download PDF
47. Motion in the interiors and atmospheres of Jupiter and Saturn: scale analysis, anelastic equations, barotropic stability criterion
- Author
-
Andrew P. Ingersoll and David Pollard
- Subjects
Convection ,Physics ,Atmosphere of Jupiter ,Equations of motion ,Astronomy and Astrophysics ,Mechanics ,Physics::Fluid Dynamics ,Classical mechanics ,Space and Planetary Science ,Inviscid flow ,Barotropic fluid ,Physics::Space Physics ,Differential rotation ,Astrophysics::Earth and Planetary Astrophysics ,Adiabatic process ,Physics::Atmospheric and Oceanic Physics ,Marginal stability - Abstract
If Jupiter's and Saturn's fluid interiors were inviscid and adiabatic, any steady zonal motion would take the form of differentially rotating cylinders concentric about the planetary axis of rotation. B. A. Smith et al. [Science215, 504–537 (1982)] showed that Saturn's observed zonal wind profile extends a significant distance below cloud base. Further extension into the interior occurs if the values of the eddy viscosity and superadiabaticity are small. We estimate these values using a scaling analysis of deep convection in the presence of differential rotation. The differential rotation inhibits the convection and reduces the effective eddy viscosity. Viscous dissipation of zonal mean kinetic energy is then within the bounds set by the internal heat source. The differential rotation increases the superadiabaticity, but not so much as to eliminate the cylindrical structure of the flow. Very large departures from adiabaticity, necessary for decoupling the atmosphere and interior, do not occur. Using our scaling analysis we develop the anelastic equations that describe motions in Jupiter's and Saturn's interiors. A simple problem is solved, that of an adiabatic fluid with a steady zonal wind varying as a function of cylindrical radius. Low zonal wavenumber perturbations are two dimensional (independent of the axial coordinate) and obey a modified barotropic stability equation. The parameter analogous to β is negative and is three to four times larger than the β for thin atmospheres. Jupiter's and Saturn's observed zonal wind profiles are close to marginal stability according to this deep sphere criterion, but are several times supercritical according to the thin atmosphere criterion.
- Published
- 1982
- Full Text
- View/download PDF
48. Carbon dioxide-water clathrate as a reservoir of CO2 on Mars
- Author
-
Anthony R. Dobrovolskis and Andrew P. Ingersoll
- Subjects
Carbon dioxide clathrate ,Martian ,Materials science ,Atmospheric pressure ,Clathrate hydrate ,chemistry.chemical_element ,Astronomy and Astrophysics ,Atmosphere of Mars ,Mars Exploration Program ,Astrobiology ,chemistry.chemical_compound ,chemistry ,Space and Planetary Science ,Carbon dioxide ,Carbon - Abstract
It has been suggested that the residual polar caps of Mars contain a reservoir of permanently frozen carbon dioxide which is controlling the atmospheric pressure. However, observational data and models of the polar heat balance suggest that the temperatures of the Martian poles are too high for solid CO_2 to survive permanently. On the other hand, the icelike compound carbon dioxide-water clathrate (CO_2·6H_2O) could function as a CO_2 reservoir instead of solid CO_2, because it is stable at higher temperatures. This paper shows that the permanent polar caps may contain several millibars of CO_2 in the form of clathrate, and discusses the implications of this permanent clathrate reservoir for the present and past atmospheric pressure on Mars.
- Published
- 1975
- Full Text
- View/download PDF
49. Merging of vortices in the atmosphere of Jupiter: An analysis of voyager images
- Author
-
Mordecai-Mark Mac Low and Andrew P. Ingersoll
- Subjects
Atmosphere ,Physics ,Jupiter ,Spots ,Space and Planetary Science ,Atmosphere of Jupiter ,Great Red Spot ,Astronomy and Astrophysics ,Tourbillon ,Astrophysics ,Shear zone ,Atmospheric sciences ,Vortex - Abstract
We have studied interactions between stable oval structures (spots) using the Voyager 2 cylindrical projection mosaics. In contrast with the solitary wave type of interaction, collisions between spots are irreversible. Most interactions (23 out of 27 cases) lead to merging of the two original spots. The other type of interaction (4 out of 27 cases) is simply a near miss—the spots pass around each other. Interactions of spots with filamentary regions, which are actively changing and more amorphous than spots, usually lead to the disappearance of the spot. Filamentary regions are also the major source of spots. Stable spots do not produce other spots. Instead, spots destroy each other by merging. Most spots are anticyclonic and sit in anticyclonic shear zones. Filamentary regions are cyclonic and sit in cyclonic shear zones. Larger spots are more elliptical than smaller ones. The most common spots have major diameters of 2000 km and minor diameters of 1500 km.
- Published
- 1986
- Full Text
- View/download PDF
50. Seasonal meridional energy balance and thermal structure of the atmosphere of Uranus: A radiative-convective-dynamical model
- Author
-
Andrew P. Ingersoll and James Friedson
- Subjects
Physics ,Convection ,Convective heat transfer ,Uranus ,Astronomy and Astrophysics ,Atmospheric sciences ,Atmosphere ,Heat flux ,Space and Planetary Science ,Physics::Space Physics ,Heat transfer ,Astrophysics::Solar and Stellar Astrophysics ,Astrophysics::Earth and Planetary Astrophysics ,Atmosphere of Uranus ,Internal heating ,Physics::Atmospheric and Oceanic Physics - Abstract
The seasonal meridional energy balance and thermal structure of the atmosphere of Uranus is investigated using a two-dimensional radiative—convective—dynamical model. Diurnal-average temperatures and heat fluxes are calculated as a function of pressure, latitude, and season. In addition to treating radiation and small-scale convection in a manner typical of conventional radiative—convective models, the dynamical heat fluxes due to large-scale baroclinic eddies are included and parametrized using a mixing length formulation (P.H. Stone, 1972; J. Atmos. Sci.29, 405–418; A. P. Ingersoll and C. C. Porco, 1978, Icarus35, 27–43). The atmosphere is assumed to be bounded below by an adiabatic, fluid interior with a single value of potential temperature at all latitudes. The internal heat flux is found to vary with latitude and season. The total internal power and the global enthalpy storage rate are seen to oscillate in phase with a period of 1/2 Uranian year. On an annual-average basis, equatorward heat transport can take place both in the atmosphere and convective interior. For a weak internal heat source, the meridional transport takes place predominantly in the atmosphere. If the internal heat source is larger, a greater share of the transport is taken up by the interior. For a value of the internal heat near the current upper limit for Uranus (∼27% of the adsorbed sunlight), about one-third of the equatorward heat transport at midlatitudes occurs in the interior. For a given internal heat source, placing the peak of the solar heating at high altitudes or depositing the solar energy into a narrow altitude range favors heat transport by the atmosphere over the interior. Deep penetration of sunlight favors transport by the interior. For the time corresponding to the Voyager 2 Uranus encounter, the effective temperature at the south (sunlit) pole is calculated to be ∼1.5°K higher than that at the equator. Horizontal contrasts of the mean 450- to 900 mbar temperature are found to be ≤1.5°K, in fair agreement with Voyager 2 IRIS results (R. Hanel et al., 1986, Science233, 70–74), but the model fails to reproduce the local minimum in this temperature seen at −30°S. Nevertheless, it is concluded that meridional heat transport in the atmosphere is efficient in keeping seasonal horizontal temperature contrasts below those predicted by radiative-convective models (L. Wallace, 1983, Icarus 54, 110–132).
- Published
- 1987
- Full Text
- View/download PDF
Catalog
Discovery Service for Jio Institute Digital Library
For full access to our library's resources, please sign in.