20.01.2015 Views

modeling jupiter's atmospheric dynamics - EPIC Atmospheric Model ...

modeling jupiter's atmospheric dynamics - EPIC Atmospheric Model ...

modeling jupiter's atmospheric dynamics - EPIC Atmospheric Model ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

MODELING JUPITER’S ATMOSPHERIC DYNAMICS<br />

WITH AN ACTIVE HYDROLOGICAL CYCLE<br />

By<br />

Csaba Palotai<br />

M.S., Technical University of Budapest, Hungary 2001<br />

A Dissertation<br />

Submitted to the Faculty of the<br />

Graduate School of the University of Louisville<br />

in Partial Fulfillment of the Requirements<br />

for the Degree of<br />

Doctor of Philosophy<br />

Department of Mechanical Engineering<br />

University of Louisville<br />

Louisville, Kentucky<br />

August 2006


MODELING JUPITER’S ATMOSPHERIC DYNAMICS<br />

WITH AN ACTIVE HYDROLOGICAL CYCLE<br />

By<br />

Csaba Palotai<br />

M.S., Technical University of Budapest, Hungary 2001<br />

A Dissertation Approved On<br />

Date<br />

by the following Dissertation Committee:<br />

Dissertation Director<br />

ii


ACKNOWLEDGEMENTS<br />

I foremost wish to express gratitude to my parents and grandparents. Sometimes,<br />

as teens or young adults we don’t realize how much people sacrifice for us, and<br />

looking back later we recognize all the great things they have done for us, even if it<br />

was hard for them or they had to give up something for us. They were also letting<br />

me go for this amazing journey in my life far away from them, which must have been<br />

the toughest thing for them to do. I will do my utmost to earn all this.<br />

I would like to thank to my advisor, Professor Tim Dowling for his help, guidance<br />

and constructive criticism, through thick and thin. He was the source of all the<br />

planetary science that I know today. He also helped through a continuous revision of<br />

this dissertation, correcting all the 35,945 grammar errors that I have made. I appreciate<br />

the helpful comments of Adam Showman, Ray LeBeau, Glenn Orton, Gerard<br />

Williger and Aaron Heirrnstein, they were very generous with sharing their time and<br />

expertise. I express my appreciation to NASA, supporting this research through their<br />

EPSCoR, Planetary Atmospheres and Outer Planet Research Programs.<br />

I would like to thank the members of my dissertation committee, Ellen Brehob,<br />

Geoffrey Cobourn, John Kielkopf, and Keith Sharp for their service and support. I<br />

wish to express my appreciation to Dr. Kielkopf for the LaTeX dissertation template<br />

that saved me a great deal of time. Several friends have been helping me along the<br />

way: Tanya, Grace, Harrison, Móni, Tomi, Marjorie, thank you for listening to (and<br />

ignoring) all my whinings and encouraging me when I needed it.<br />

I would like to<br />

iii


acknowledge all my great teachers throughout the years, I have been very fortunate<br />

because there were so many of them. Thank you for all your lessons about life.<br />

And most of all, I would like to thank my family. I have been working for<br />

them, but they had to sacrifice a lot too. Benita, thank you for all your patience and<br />

for accepting me for who I am. You are my Lobstress. Junior, I am so happy that<br />

you have arrived. You are so small yet, but you already have the power to fill our<br />

hearts with happiness. I am looking forward to all the joy and great times that we<br />

will share. And Ben, your energy and love have been so inspiring. Every day is an<br />

adventure with you and I have been having the time of my life. You make me want<br />

to be a better parent and a better person every day. I’m eager to find out where the<br />

next steps in our journey will take us.<br />

iv


ABSTRACT<br />

MODELING JUPITER’S ATMOSPHERIC DYNAMICS WITH AN ACTIVE<br />

HYDROLOGICAL CYCLE<br />

Csaba Palotai<br />

June 1, 2006<br />

An active hydrological cycle has been added to the <strong>EPIC</strong> general circulation<br />

model (GCM) for planetary applications, with a special emphasis on Jupiter. Scientists<br />

have suspected for decades that clouds, and in particular latent heating, strongly<br />

influence Jupiter’s <strong>atmospheric</strong> <strong>dynamics</strong> and this research provides a tool to investigate<br />

this phenomenon. Components of the model have been adapted for the planetary<br />

setting from recently published Earth microphysics schemes.<br />

The behavior of the cloud model is investigated in two steps. First, we explore<br />

in detail the runtime properties of a nominal model, and second, through sensitivity<br />

tests we determine how the full microphysics and selected components of the scheme<br />

affect the formation and evolution of clouds and precipitation.<br />

Results from our<br />

one-dimensional (vertical) simulations match expectations based on thermochemical<br />

models about the vertical positioning of ammonia and water clouds, and the nature of<br />

precipitation. Using meridional-plane (two-dimensional) simulations, we investigate<br />

the latitudinal variation of clouds. We conclude that the the zonal-wind structure<br />

vi


under the visible cloud deck strongly affects the position of the cloud bases.<br />

We<br />

describe in detail an equatorial storm system observed in our 2D simulations. We<br />

also show that simplification of our microphysics scheme would improperly simulate<br />

large scale weather phenomena on Jupiter.<br />

We support future laboratory tests and in-situ measurements that would improve<br />

the cloud parameterization scheme and would also add more constraints on the<br />

global distribution of condensibles and on the zonal wind-structure. The complete<br />

computer program resulting from this research can be downloaded as open-source<br />

software from NASA’s Planetary Data System (PDS) Atmospheres node.<br />

vii


TABLE OF CONTENTS<br />

Page<br />

ACKNOWLEDGEMENTS<br />

ABSTRACT<br />

LIST OF TABLES<br />

LIST OF FIGURES<br />

iii<br />

vi<br />

xi<br />

xii<br />

CHAPTER<br />

1 BACKGROUND AND MOTIVATION 1<br />

1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1<br />

1.2 The <strong>EPIC</strong> <strong>Model</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . 4<br />

1.3 Review of Previous Studies . . . . . . . . . . . . . . . . . . . . 6<br />

1.3.1 Published <strong>EPIC</strong> Simulations . . . . . . . . . . . . . . . . 6<br />

1.3.2 Earth <strong>Model</strong>s . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

1.3.3 Planetary <strong>Atmospheric</strong> <strong>Model</strong>s . . . . . . . . . . . . . . . 9<br />

1.4 Logistics for Experiments . . . . . . . . . . . . . . . . . . . . . 10<br />

1.4.1 Simulation Tools . . . . . . . . . . . . . . . . . . . . . . 11<br />

1.4.2 Hardware Tools . . . . . . . . . . . . . . . . . . . . . . . 11<br />

1.4.3 Analysis Tools . . . . . . . . . . . . . . . . . . . . . . . . 12<br />

1.5 Scope of Research . . . . . . . . . . . . . . . . . . . . . . . . . . 12<br />

2 MICROPHYSICS SCHEME 14<br />

2.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14<br />

viii


2.2 Moisture Equations . . . . . . . . . . . . . . . . . . . . . . . . . 16<br />

2.3 Number Concentration of Ice Crystals . . . . . . . . . . . . . . 17<br />

2.4 Precipitation Size Distribution . . . . . . . . . . . . . . . . . . . 22<br />

2.5 Fall Speed of Particles . . . . . . . . . . . . . . . . . . . . . . . 25<br />

2.5.1 Raindrops . . . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

2.5.2 Terminal Velocity of Snow . . . . . . . . . . . . . . . . . 29<br />

2.6 Microphysical Processes . . . . . . . . . . . . . . . . . . . . . . 33<br />

2.6.1 Condensation of Vapor/Evaporation of Cloud Liquid (P COND ) 34<br />

2.6.2 Initiation of Cloud Ice (P INIT ) . . . . . . . . . . . . . . . 36<br />

2.6.3 Deposition of Vapor/Sublimation of Cloud Ice (P DEPI ) . 38<br />

2.6.4 Melting of Cloud Ice/Freezing of Cloud Liquid and<br />

Melting of Snow/Freezing of Rain (P SMLTI ) . . . . . . . 40<br />

2.6.5 Autoconversion of Cloud Liquid to Rain (P RAUT ) . . . . 41<br />

2.6.6 Collection of Cloud Liquid by Rain (P RACW ) . . . . . . 41<br />

2.6.7 Deposition of Vapor on Rain/Evaporation of Rain (P REVP ) 42<br />

2.6.8 Autoconversion of Cloud Ice to Snow (P SAUT ) . . . . . . 43<br />

2.6.9 Collection of Cloud Ice by Snow (P SACI ) . . . . . . . . . 44<br />

2.6.10 Deposition of Vapor on Snow/Sublimation of Snow (P SEVP ) 44<br />

2.7 Code Implementation . . . . . . . . . . . . . . . . . . . . . . . 45<br />

2.7.1 Positivity adjustment . . . . . . . . . . . . . . . . . . . . 45<br />

2.7.2 Sequence of Process Transfers . . . . . . . . . . . . . . . 46<br />

3 1D SIMULATIONS 48<br />

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48<br />

3.2 Nominal <strong>Model</strong> Description . . . . . . . . . . . . . . . . . . . . 49<br />

ix


3.2.1 Vertical Structure of <strong>Model</strong> Environment . . . . . . . . . 49<br />

3.2.2 Vertical Structure of Clouds . . . . . . . . . . . . . . . . 51<br />

3.3 Results of 1D Simulations . . . . . . . . . . . . . . . . . . . . . 54<br />

3.3.1 Sensitivity to abundance . . . . . . . . . . . . . . . . . . 57<br />

3.3.2 Static Stability of the Atmosphere . . . . . . . . . . . . . 61<br />

4 MERIDIONAL PLANE SIMULATIONS 65<br />

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65<br />

4.2 <strong>Model</strong> Description . . . . . . . . . . . . . . . . . . . . . . . . . 66<br />

4.2.1 Vertical Structure of <strong>Model</strong> Environment . . . . . . . . . 66<br />

4.2.2 Vertical Structure of Condensibles . . . . . . . . . . . . . 68<br />

4.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68<br />

4.3.1 Equatorial Storm System . . . . . . . . . . . . . . . . . . 71<br />

4.3.2 Sensitivity tests . . . . . . . . . . . . . . . . . . . . . . . 77<br />

5 CONCLUSIONS 85<br />

REFERENCES 89<br />

CURRICULUM VITAE 99<br />

x


LIST OF TABLES<br />

TABLE<br />

Page<br />

2.1 Saturation vapor pressure coefficients. . . . . . . . . . . . . . . . . . . 18<br />

2.2 Power law coefficients used in the <strong>EPIC</strong> microphysics code to calculate<br />

N ice for water and ammonia (2.6–2.11). . . . . . . . . . . . . . . . . . 23<br />

3.1 Microphysical components turned on (+) and off (-) in the Brunt-<br />

Väisälä test runs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62<br />

xi


LIST OF FIGURES<br />

FIGURE<br />

Page<br />

1.1 Jupiter’s Great Red Spot and its turbulent zone to the west. . . . . . 2<br />

1.2 A jovian storm, color-coded to highlight structure. . . . . . . . . . . . 3<br />

1.3 Comparison of visual image of methane clouds on Neptune with <strong>EPIC</strong><br />

simulations using a diagnostic (passive) cloud scheme. . . . . . . . . . 7<br />

2.1 Correlation between the Reynolds- and Best-numbers in the Bestnumber<br />

regime of 10 3 to 10 5 . . . . . . . . . . . . . . . . . . . . . . . . 29<br />

2.2 Comparison of terminal velocities of raindrops and snow. . . . . . . . 30<br />

2.3 Schematic diagram of the eleven phase-change processes included in<br />

the five-phase <strong>EPIC</strong> microphysics model. . . . . . . . . . . . . . . . . 35<br />

2.4 Comparison of initialized ice mass calculated from (2.5) for water on<br />

Earth (solid line), and from (2.40) for ammonia (dash-dot line) and<br />

water (dotted line) on Jupiter. . . . . . . . . . . . . . . . . . . . . . . 38<br />

3.1 Basic state temperature and mean zonal wind profiles used in the 1D<br />

simulations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50<br />

3.2 <strong>Model</strong>s of the vertical profiles of ammonia and water volume mixing<br />

ratios used in this study. . . . . . . . . . . . . . . . . . . . . . . . . . 53<br />

3.3 The time evolution of water and ammonia clouds and precipitation in<br />

the nominal 1D model of Jupiter’s atmosphere. . . . . . . . . . . . . 56<br />

3.4 Cloud and precipitation evolution using various initial deep abundance<br />

values, with initial supersaturation set to 10%. . . . . . . . . . . . . 59<br />

xii


3.5 Sensitivity of cloud and precipitation evolution to various initial supersaturation<br />

values, with the deep abundance values set to the nominal<br />

case of 4.0 and 1.0 times solar for ammonia and water, respectively. . 60<br />

3.6 Brunt-Väisälä frequency of the initial state. . . . . . . . . . . . . . . 62<br />

3.7 Brunt-Väisälä frequency at 2 hours into the simulation with selected<br />

microphysics components turned off. . . . . . . . . . . . . . . . . . . 64<br />

4.1 Initial state temperature and mean zonal wind profiles used in the 2D<br />

simulations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67<br />

4.2 Vertical structure of clouds in the nominal 2D model. . . . . . . . . . 69<br />

4.3 emporal evolution of the latitudinal variation of cloud density for ammonia<br />

(top) and water (bottom), sampled at the 750 mbar and the<br />

4000 mbar pressure levels, respectively. . . . . . . . . . . . . . . . . . 72<br />

4.4 Time evolution of the equatorial storm in the nominal model. . . . . 73<br />

4.5 40 day difference of zonal mean temperature compared to the initial<br />

state of the active simulation. . . . . . . . . . . . . . . . . . . . . . . 74<br />

4.6 Close up images of the equatorial storm. Panels on the left show clouds<br />

and wind velocity vectors, panels on the right show clouds and precipitation.<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76<br />

4.7 Zonal cloud structure with modified initial saturation ratio values. . . 78<br />

4.8 Cloud and zonal wind structure at 30 days into the active simulation<br />

using S = 1.00 initial saturation ratio. . . . . . . . . . . . . . . . . . . 80<br />

4.9 Sensitivity test in the domain of -20 ◦ to 60 ◦ latitude. . . . . . . . . . 80<br />

4.10 Test results with reduced abundance initial condition . . . . . . . . . 81<br />

4.11 Snapshots from a simulation, where water was the only condensible<br />

species. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83<br />

xiii


4.12 Cloud structure in model runs with selected microphysical components<br />

turned off, at 40 days into the active simulation. . . . . . . . . . . . 83<br />

xiv


CHAPTER 1<br />

BACKGROUND AND MOTIVATION<br />

1.1 Motivation<br />

The well-known features of Jupiter’s atmosphere, the Great Red Spot (GRS,<br />

Fig. 1.1), and the dark and bright bands called belts and zones, have been observed<br />

for several centuries, yet the energy sources that shape the active meteorology on<br />

the planet, forming and maintaining its long lived vortices and jets, are not well understood.<br />

Recent results from the Galileo spacecraft answer several questions about<br />

Jupiter’s energetics, but leave many others unanswered. One of the most promising<br />

candidates for driving the motions is the internal heat source (Gierasch et al. 2000,<br />

Seiff 2000). Observations of lightning (Little et al. 1999) imply precipitation in convective<br />

storms, thus these observations indicate moist convection on Jupiter. These<br />

storms (Fig. 1.2) resemble clusters of thunderstorm cells on Earth called mesoscale<br />

convective complexes (MCCs).<br />

On our planet, the atmosphere and the surface are strongly influenced by the<br />

clouds and precipitation through phenomena like their transport of heat, moisture,<br />

and aerosols, the latent heat from phase changes of water, the alteration of the <strong>atmospheric</strong><br />

and surface radiative fluxes, and their effect on surface hydrology. For a more<br />

accurate terrestrial weather and climate simulation one needs to take these effects<br />

into account when using <strong>atmospheric</strong> models. The software component that handles<br />

1


Figure 1.1. Jupiter’s Great Red Spot and its turbulent zone to the west. The<br />

erupting white clouds northwest of the GRS occur relatively often in this region, as<br />

was shown by Galileo and other data (Image:NASA/Voyager1)<br />

2


Figure 1.2. A jovian storm, color-coded to highlight structure. Light areas indicate<br />

high haze. Deep red in the lower right shows storm clouds. Researchers suspect the<br />

clouds result from condensed water driven by convection. The schematic at right is<br />

a cross-section of the storm showing the cloud structure. This is an example of the<br />

phenomena we wish to begin to capture with the cloud microphysics scheme developed<br />

here. (Image: P. Gierasch/ c○Nature)<br />

the formation and evolution of clouds is usually referred to as the “cloud microphysics<br />

model”.<br />

General circulation models (GCMs) make it possible to simulate an atmosphere<br />

realistically without using any artificial boundary conditions.<br />

GCMs allow<br />

one to investigate idealized or localized processes to improve our knowledge about<br />

the underlying <strong>dynamics</strong>. They serve as a testing tool for verifying new ideas, also<br />

to make clear what observational input is most sensitive to a particular process. The<br />

GCM developed at University of Louisville is called the Explicit Planetary Isentropic<br />

Coordinate (<strong>EPIC</strong>) <strong>atmospheric</strong> model (Dowling et al. 1998, Dowling et al. 2006).<br />

It is specifically designed for comparative-planetology studies, with the potential to<br />

simulate all the dozen or so atmospheres in the Solar System with a single model. It is<br />

the leading software for investigating the <strong>atmospheric</strong> <strong>dynamics</strong> and thermo<strong>dynamics</strong><br />

of gas-giant planets.<br />

One major component missing from the <strong>EPIC</strong> simulations thus far is the effect<br />

3


of cloud microphysical processes. There has been one <strong>EPIC</strong> paper involving clouds to<br />

date, a study of methane cloud formation on Neptune (Stratman et al. 2001). They<br />

diagnosed clouds when the relative humidity of methane reached 100% or greater,<br />

assumed non-precipitating clouds, and neglected the effects of latent heating. This<br />

approach can be justified for certain applications when the abundance of the condensible<br />

species is low or the latent heat of phase change is low. In other applications<br />

(e.g. water clouds on Earth or on Jupiter) latent heating plays a significant role and<br />

consequently needs to be accounted for (Ingersoll et al. 2004). The objective of this<br />

research is the addition of a cloud microphysics scheme to the <strong>EPIC</strong> <strong>atmospheric</strong><br />

model, which allows for the simulation of the full hydrological cycle. While the particular<br />

emphasis is on Jupiter, the cloud microphysics scheme is flexible and can be<br />

easily adapted to other planetary atmospheres.<br />

1.2 The <strong>EPIC</strong> <strong>Model</strong><br />

The governing equations of the <strong>EPIC</strong> model are the hydrostatic primitive equations<br />

in spherical coordinates with a hybrid vertical coordinate, similar to those used<br />

by Konor and Arakawa (1997, hereafter KA97). This vertical coordinate, ζ, uses the<br />

advantages of an isentropic θ-coordinate near the top of the model that transitions<br />

smoothly into a terrain following pressure-based coordinate, σ (Dowling et al. 2006).<br />

In the case of gas giant planets, which do not have solid lower boundaries but have<br />

a nearly constant θ interior, the transition is into a pressure coordinate that allows<br />

one to model the deeper atmosphere. The material time derivative with the hybrid<br />

coordinate can be expressed as<br />

( )<br />

D ∂<br />

Dt = + v · ∇ ζ +<br />

∂t<br />

˙ζ ∂ ∂ζ , (1.1)<br />

ζ<br />

4


where ˙ζ ≡ Dζ/Dt, ∇ ζ is the horizontal gradient on constant ζ surfaces, and v is the<br />

horizontal velocity on the ζ surfaces. One can introduce the variable h, defined by<br />

h ≡ − 1 g<br />

∂p<br />

∂ζ , (1.2)<br />

which for a shallow (hydrostatic) atmosphere plays the role of density in the (x, y, ζ)<br />

space, where p is pressure and g is the gravitational acceleration (assumed constant).<br />

In the case of a dry atmosphere (i.e. no condensible species, only a homogeneous gas<br />

mixture) using the hybrid density h, the continuity equation may be written as<br />

( ) ∂h<br />

+ ∇ ζ · (hv) + ∂ ∂t<br />

∂ζ (h ˙ζ) = SHS h + SZF h , (1.3)<br />

ζ<br />

where SHS h and SZF h represent the rates of change due to the horizontal eight order<br />

smoothing (hyperviscosity) and the high latitude zonal filtering, respectively (Dowling<br />

et al. 2006). In this coordinate system the hydrostatic equation may be written in<br />

the form of<br />

∂Φ<br />

∂ζ<br />

= αgh, (1.4)<br />

where α is the specific volume, and Φ is the geopotential, defined by<br />

where z is the geometric height.<br />

is given by<br />

Φ ≡<br />

∫ z<br />

0<br />

gdz, (1.5)<br />

We assume the model atmosphere to be an ideal gas, thus the equation of state<br />

α = RT<br />

p , (1.6)<br />

where R is the gas constant and T is the temperature. Dowling et al. (1998) define<br />

the mean potential temperature, θ, for Jupiter taking into account for the ortho-para<br />

hydrogen mixture. The 1st Law of thermo<strong>dynamics</strong> can thus be written as<br />

( )<br />

Dθ ∂θ<br />

Dt = ∂θ<br />

+ v · ∇ ζ θ + ˙ζ<br />

∂t<br />

∂ζ = ˙Q<br />

Π + SHS θ + SZF θ + STB θ , (1.7)<br />

ζ<br />

5


where ˙Q is the heating rate per unit mass, STB θ represents the rate of change of θ<br />

due to the turbulent diffusion, and Π is the Exner function, defined by<br />

Π = c pT<br />

θ , (1.8)<br />

where c p is the specific heat at constant pressure. The horizontal momentum equation<br />

may be written as<br />

( )<br />

Dv ∂v<br />

Dt = + v · ∇ ζ v +<br />

∂t<br />

ζ<br />

˙ζ<br />

∂v<br />

∂ζ = −(∇ pΦ) − fk × v + F, (1.9)<br />

where ∇ p is the horizontal gradient on constant pressure surfaces, F is the friction<br />

force, k is the vertical unit vector, f = 2Ω sin λ is the Coriolis parameter, Ω is the<br />

planet’s angular velocity, and λ is the latitude. After combining equations similarly<br />

to KA97, we can express the horizontal pressure gradient force as<br />

−∇ p Φ = −∇ ζ M + Π∇ ζ θ, (1.10)<br />

where M is the Montgomery potential, defined by<br />

M ≡ c p T + Φ. (1.11)<br />

1.3 Review of Previous Studies<br />

1.3.1 Published <strong>EPIC</strong> Simulations<br />

There are several previously published results using the <strong>EPIC</strong> model. For example,<br />

simulations of the impact of Comet Shoemaker-Levy 9 (SL9) on Jupiter that<br />

successfully predicted that there would be no long-lasting vortices spawned from that<br />

event (Harrington et al.<br />

1994), followed by simulations of time-dependent, threedimensional<br />

vortices under conditions similar to those on Neptune that showed relation<br />

between the meridional drift of the vortices and the background absolute vorticity<br />

6


Figure 1.3. Comparison of visual image of methane clouds on Neptune with <strong>EPIC</strong><br />

simulations using a diagnostic (passive) cloud scheme. Isosurfaces of 100% relative<br />

humidity are white and of eddy relative vorticity of 8.0 × 10 −7 s −1 are blue. The<br />

bottom row is a side view looking north (Image: NASA)<br />

(LeBeau and Dowling 1998). The latter study also investigated oscillations in aspect<br />

ratio and orientation angle, and tail formation, and motivated a three-dimensional<br />

simulation of the bright companion clouds associated with Neptune’s Great Dark<br />

Spot that predicted methane clouds at or just below the planet’s tropopause using a<br />

diagnostic cloud scheme (Stratman et al. 2001, Figure 1.3). In addition, Showman<br />

and Dowling (2000) modeled Jupiter’s 5µm hot spots to simulate the dryness and low<br />

cloudiness measured by the Galileo probe, and recently, Garcia-Melendo et al. (2005)<br />

tested various vertical temperature and wind profiles in order to reproduce observed<br />

disturbances in Jupiter’s highest velocity jet at 24 ◦ N latitude. None of the published<br />

<strong>EPIC</strong> simulations to date include an active hydrological cycle, which is the main new<br />

functionality introduced in this work.<br />

7


1.3.2 Earth <strong>Model</strong>s<br />

Recently, terrestrial-atmosphere GCM modelers have paid more and more attention<br />

to the development of cloud parameterizations and the inclusion of interactions<br />

between the hydrological cycle, cloudiness, and radiation in terms of cloud microphysical<br />

processes and radiative properties. <strong>Model</strong>ing clouds in large-scale models<br />

is a complex problem. There are several different cloud types, and their precipitation<br />

content, latent heating, optical depth, and some of their microphysical processes vary<br />

from one type to another. They are formed, maintained and dissipated by several different<br />

physical processes, such as convection, large-scale ascent or descent, small-scale<br />

turbulence, and cloud microphysical processes leading to the generation of precipitation<br />

of different forms. Moreover, different cloud types may coexist within the volume<br />

represented by a single GCM scale grid-cell.<br />

Fowler et al.<br />

(1996) gives an overview on how clouds are typically represented<br />

in large-scale models. Early schemes used a diagnostic approach where the<br />

clouds were predicted based on a prescribed threshold value of relative humidity for<br />

saturation, and cloud fraction and optical properties were diagnosed from selected<br />

large-scale variables, such as relative humidity, rate of precipitation, temperature,<br />

etc. A diagnostic, saturation adjustment cloud scheme is a component of the earlier<br />

<strong>EPIC</strong> version, but its limitations for <strong>modeling</strong> jovian meteorology were experienced<br />

by Stratman et al. (2001).<br />

Some other prognostic schemes employ a “double-moment” parameterization<br />

in which both number concentration of particles and mass mixing ratio are predicted<br />

at each grid location (e.g. Ferrier 1994). This method allows one to simulate the<br />

evolution of the hydrometeor size spectra more realistically. However, at this point<br />

we have ruled out using this type of <strong>modeling</strong> because of the lack of knowledge of the<br />

8


number concentration of the various species on Jupiter.<br />

A more widely used approach for representing clouds and precipitation in<br />

GCMs is a bulk parameterization scheme where the number concentrations are diagnosed<br />

instead of being predicted. Nowadays, almost all terrestrial models use prognostic<br />

equations for the various forms of water substance (vapor, cloud liquid, cloud<br />

ice, rain, snow, graupel, etc.), which are coupled together by the cloud microphysics.<br />

Most of the microphysical processes have smaller characteristic length scales than a<br />

typical grid length used in large-scale simulations, and many of the processes are not<br />

completely understood. For these reasons, GCMs and mesoscale models use parameterizations<br />

for most of the cloud processes rather than explicitly resolving them.<br />

1.3.3 Planetary <strong>Atmospheric</strong> <strong>Model</strong>s<br />

Many numerical cloud models exist for solid-surface planets (e.g.<br />

James et<br />

al. 1997, Forget et al. 1998, Richardson et al. 2002), and most of them use a<br />

simple cloud scheme or adopt a terrestrial scheme with minor modifications. Some<br />

of these atmospheres (e.g. Mars) tend not to precipitate, which results in a major<br />

simplification in the hydrological cycle <strong>modeling</strong>.<br />

There are some published cloud models for gas-giant planets as well.<br />

The<br />

absence of a solid surface and the lack of in situ measurements make it challenging<br />

to build an accurate model, and also make it necessary to use assumptions and neglect<br />

potentially important effects, like the feedback from the <strong>atmospheric</strong> circulation.<br />

Early jovian simulations use equilibrium cloud condensation models to study the vertical<br />

structure of clouds or chemical composition, predicting three major cloud layers<br />

for the jovian troposphere formed by particles of NH 3 , NH 4 SH, and H 2 O (Lewis 1969,<br />

Weidenschilling and Lewis 1973). Carlson et al. (1988) estimated the timescales of<br />

9


microphysical processes on Jupiter using a model with no spatial dimensions. Stoker<br />

(1986) simulated jovian equatorial plumes and cumulus cloud formation with a onedimensional<br />

diagnostic model. Her investigation regarding the importance of water<br />

and ammonia condensation produced results suggesting strong convection and high<br />

vertical velocities that extend several scale heights from the water cloud base to above<br />

the top of ammonia clouds. Lunine and Hunten (1987) applied the same model for<br />

narrow, moist convective plumes to simulate the apparent depletion of water vapor<br />

in the 5-µm observations in the lower troposphere without invoking a depletion of<br />

oxygen from solar abundance throughout the interior.<br />

Del Genio and McGratten<br />

(1990) used a GCM cumulus parameterization scheme in one dimension to study<br />

the effects of moist convection on Jupiter’s large scale <strong>atmospheric</strong> structure. Some<br />

other simulations included detailed microphysics and an axisymmetric cloud model<br />

to study jovian thunderclouds (Yair et al. 1995) or a three-dimensional model with<br />

simplified microphysics to investigate the formation and evolution of moist convective<br />

storms (Hueso and Sánchez-Lavega 2001). These are models on the regional scale<br />

and exclude the interaction with planetary size jets and vortices. There is a need for<br />

three-dimensional GCMs for gas giant planets that includes cloud microphysics to<br />

expand our knowledge on their atmospheres.<br />

1.4 Logistics for Experiments<br />

Here we mention some technical details about the software and hardware tools<br />

available for our project, and briefly describe techniques used for data visualization<br />

and analysis.<br />

10


1.4.1 Simulation Tools<br />

There are some major upgrades for the current <strong>EPIC</strong> 4.0 version that help us<br />

with this research. The new hybrid vertical coordinate system made it possible to<br />

model clouds in the deeper atmosphere, such as the water clouds on Jupiter with a<br />

cloud base at around 5 bars. Separate continuity equations are invoked for each phase<br />

of each species, with the possibility of turning any species and phases on and off, or<br />

switching the microphysics (i.e. the interaction between the phases) on and off, as<br />

needed. This tool can help to simulate the effect of latent heating on the <strong>dynamics</strong>,<br />

for example. A species variable structure has been introduced that contains physical<br />

constants and functions for each species, such as the triple-point temperature and<br />

pressure, saturation vapor pressure, enthaply change due to phase change, coefficients<br />

in the power-law form of the terminal velocity, etc.<br />

1.4.2 Hardware Tools<br />

The <strong>EPIC</strong> model uses the Message Passing Interface (MPI) standard, which<br />

enables the user to run the code on parallel computers, in order to be able to solve all<br />

these equations on a sufficiently fine mesh. For our project we built a dedicated 40-<br />

node parallel computer with employing a Flat Neighborhood Network (FNN) called<br />

the Comparative Planetology Linux Engine, COMPLINE. The server machines use<br />

the cAos Linux operational system and the whole cluster gives a 33.9 Gflop peak<br />

performance, enabling one to run full-globe climate simulations in a reasonable time.<br />

For example, a 30-day simulation, using 20 layers and a 128x256 grid, including two<br />

species with all five phases and active microphysics runs for approximately 24 hrs<br />

using 8 computing nodes.<br />

11


1.4.3 Analysis Tools<br />

Several applications have been used for analyzing the <strong>EPIC</strong> computer code<br />

and the output data. For runtime analysis, including memory corruption detection,<br />

memory leak detection, cache-miss and performance profiling, we used PurifyPlus,<br />

Valgrind and Cachegrind on the Linux servers, which helped to write faster and more<br />

reliable codes. For data analysis and visualization, we used Macintosh workstations<br />

with IDL 6.0 (Interactive Data Language), Kaleidagraph 3.6, Microsoft Excel and<br />

various Adobe applications for plotting and animation.<br />

1.5 Scope of Research<br />

We have researched several possibilities for representing the clouds and microphysical<br />

processes in the <strong>EPIC</strong> model and adapted the ones that we found relevant<br />

for our purposes (summarized by Fig. 2.3). Chapter 2 describes the details of the<br />

microphysics scheme used in the <strong>EPIC</strong> model. Since there is a lack of knowledge<br />

about several physical properties of planetary atmospheres (Rossow 1978), we have<br />

worked on simplifying and fine tuning the parameterization to be suitable for this<br />

research. For example, ammonia clouds form on Jupiter but not on Earth — how<br />

do they behave with regard to cloud fraction, particle size, or precipitation Are the<br />

clouds and precipitation processes the same in Jupiter’s hydrogen atmosphere as in<br />

Earth’s nitrogen atmosphere The answer is very likely no for both questions. If<br />

not, how should one represent the processes to be suitable for the several different<br />

atmospheres in the Solar System We used a simplified, one-dimensional version of<br />

<strong>EPIC</strong> to test and to improve the parameterizations in the microphysics scheme, and<br />

to carry out sensitivity tests for the most important unknown parameters. We present<br />

12


esults from these simulations in Chapter 3.<br />

Two-dimensional experiments (latitude-height) were used to formulate comparison<br />

between dry and moist simulations, to estimate the position and thickness<br />

of model water and ammonia clouds, and to determine the direction of vertical motion<br />

relative to the zonal-wind (or belt-zone) structure. Findings from these models<br />

are discussed in Chapter 4. Conclusions and a discussion of future work is given in<br />

Chapter 5.<br />

13


CHAPTER 2<br />

MICROPHYSICS SCHEME<br />

2.1 Overview<br />

There are several different approaches to cloud microphysics in the meteorological<br />

literature, and as we have worked to add this functionality to the <strong>EPIC</strong> model<br />

we have had to make choices at every step, some with greater certainty than others.<br />

Here, we document the most significant decisions in a manner that is intended<br />

to make it easy to revisit them as necessary, which we expect to do periodically as<br />

we gain experience with <strong>modeling</strong> clouds on gas giants. We started the process by<br />

making a comprehensive survey of published cloud microphysics models, looking for<br />

those that would be most flexible given the multi-planet nature of <strong>EPIC</strong>. We were<br />

not overly concerned about memory requirements because of the large capacity of<br />

modern computer clusters. So, for example, we tended to favor those models that<br />

treat snow and rain separately rather than having them be combined into a single<br />

‘precip’ variable, since then there are fewer artificial parameters to consider. On the<br />

other hand, given that we do not yet have strong constraints on many aspects of<br />

cloud microphysics such as the sizes and shapes of cloud particles or of snow flakes,<br />

those models that average over a range of sizes were deemed most appropriate.<br />

The net result of our literature searching has been the identification of a set of<br />

papers that we used to guide our model’s development. Most of our code is adapted<br />

14


from models that we found straightforward to accurately modify for planetary applications,<br />

such as the cloud scheme from the National Centers for Environmental Prediction’s<br />

Regional Spectral <strong>Model</strong> (NCEP RSM), which implements the algorithms of<br />

Rutledge and Hobbs (1983, hereafter RH83) and Dudhia (1989, hereafter D89), and<br />

is described by Hong et al. (2004, hereafter HDC04). We also found the microphysics<br />

scheme of the Colorado State University General Circulation <strong>Model</strong> (CSU GCM) very<br />

useful, as described by Fowler et al. (1996, hereafter FRR96).<br />

Because the <strong>EPIC</strong> model is intended to be applied to many different planetary<br />

systems, we need the flexibility of activating either zero, one, or more than one<br />

condensible species. At initialization, the user chooses from a list which to activate,<br />

and for each species chosen, a continuity equation is added to the model for each<br />

corresponding phase. Our model contains five phases for each species: vapor, liquid<br />

cloud, ice cloud, rain, and snow. We recognize that under many conditions cloud ice<br />

and cloud liquid do not coexist, and the same holds for rain and snow. However,<br />

treating them separately will allow for the straightforward <strong>modeling</strong> of mixed-phase<br />

clouds in the future, which can play an important role on Earth and may prove to be<br />

interesting on Jupiter, for example as a means of improving the <strong>modeling</strong> of chemical<br />

processes or the prediction of lightning.<br />

The rest of this chapter is organized as follows. The moist continuity equations<br />

are introduced in Section 2.2. The topics of diagnosing cloud-particle number density<br />

and size distribution are addressed in Sections 2.3 and 2.4, respectively. Terminal<br />

velocity equations for precipitation, with particular emphasis on Jupiter, are derived<br />

in Section 2.5. Microphysical interactions among the phases are described in Section<br />

2.6, and some important technical details are included in Section 2.7.<br />

15


2.2 Moisture Equations<br />

We denote the mass mixing ratio as<br />

q s,p = ρ s,p<br />

ρ dry<br />

= h s,p<br />

h dry<br />

, (2.1)<br />

where ρ refers to conventional density in units of kg m −3 , h refers to the hybrid density<br />

in units of kg m −2 K −1 , which is the density associated with the hybrid-coordinate volume<br />

element, the subscripts “s, p” indicate the species and phase, and the subscript<br />

“dry” denotes the dry-air value. In the core <strong>EPIC</strong> model, the prognostic variable<br />

for moisture content is the hybrid density, h s,p = h dry q s,p , whereas in the cloud microphysics<br />

subroutines we deal primarily with the mixing ratios themselves, q s,p . In<br />

the following discussion, if the species is unambiguous we drop the species subscript<br />

and refer simply to the vapor, liquid cloud, ice cloud, rain, and snow phases as q vap ,<br />

q liq , q ice , q rain and q snow , respectively. Relative humidity is calculated according to the<br />

definition of the World Meteorological Organization as the ratio of the vapor mass<br />

mixing ratio, q vap , to the saturation mixing ratio, q sat (see Bohren and Albrecht 1998<br />

for an interesting discussion concerning this definition)<br />

r = q vap<br />

q sat<br />

=<br />

q vap<br />

ɛe sat /p dry<br />

, (2.2)<br />

where ɛ is the ratio of the dry air and vapor gas constants, p dry is the pressure of dry<br />

air, e sat is the saturation vapor pressure, and the ideal-gas model is assumed. For each<br />

species the saturation vapor pressures over liquid and ice are calculated assuming the<br />

form<br />

log 10 e sat = a + b T , (2.3)<br />

where T is the temperature in degree Kelvin, and the coefficients a and b are taken<br />

from Lodders and Fegley (1998) and listed in Table 2.1.<br />

16


Inclusion of moisture in the model introduces new continuity equations<br />

∂<br />

∂t (h dryq s,p ) + ∇ ζ · (h dry q s,p v) + ∂ ∂ζ (h dryq s,p ˙ζ) = hdry Ŝ s,p , (2.4)<br />

where Ŝs,p represents the net mixing-ratio source/sink rate for each phase of each<br />

species. The heart of a cloud-microphysics scheme is the manner in which it evaluates<br />

its Ŝs,p terms, so we include a detailed discussion of our calculations below.<br />

With clouds come falling snow and rain, and there are various levels of approximation<br />

one can use to account for the relative motion of precipitation. One issue is<br />

geometric—the hybrid-coordinate surfaces are not geopotential surfaces and so there<br />

is in general a horizontal component to the falling velocity of snow and rain. At this<br />

point, we have not taken this effect into account; we argue that it is relatively small<br />

and comparable to the errors in our estimation of the terminal fall speeds. We also<br />

currently do not include any drag on the atmosphere from the falling precipitation,<br />

but we plan to consider this in the future since it may have a small impact on simulations.<br />

To account for the relative fall speed of precipitation, the hybrid vertical<br />

velocity, ˙ζ, in (2.4) is modified by −|Vt |ρ/h, where |V t | is the terminal fall speed in<br />

ms −1 , the calculation of which is described below, and ρ/h = dζ/dz is the ratio of<br />

the full (moist air) densities. The other major change to the model caused by the<br />

inclusion of moisture is the addition of latent heating and cooling to the heating rate,<br />

˙Q, which is straightforward.<br />

2.3 Number Concentration of Ice Crystals<br />

We start with a description of the models and their parameters that are needed<br />

to characterize the ice and liquid particles that make up clouds and precipitation,<br />

given that they are much too small to be explicitly resolved by our GCM grid. In<br />

17


TABLE 2.1<br />

Saturation vapor pressure coefficients used in (2.3). The subscripts “ice” and “liq”<br />

refer to whether the saturation is over solid or liquid material, as determined by the<br />

temperature. Values are from Table 1.20 of Lodders and Fegley (1998), with 5 added<br />

to the a columns to yield e sat in Pa instead of bar.<br />

Species a ice b ice a liq b liq<br />

H 2 O 12.610 -2681.18 11.079 -2261.1<br />

NH 3 11.900 -1588.00 10.201 -1248.0<br />

H 2 S 10.610 -1171.20 9.780 -1015.5<br />

CH 4 9.283 -475.60 9.092 -459.8<br />

CO 2 12.025 -1336.00 11.045 -1201.0<br />

NH 4 SH 12.600 -2411.20 n/a n/a<br />

O 3 n/a n/a 8.912 -632.4<br />

N 2 9.798 -360.20 8.944 -305.0<br />

this project we consider both water (H 2 O) and ammonia (NH 3 ), but for simplicity we<br />

exclude H 2 S and its corresponding NH 4 SH cloud, which is thought to form between<br />

Jupiter’s water and ammonia clouds, to avoid for now the additional complication<br />

of active chemistry. Some characteristics of individual particles are sensitive to the<br />

details, but if we average over an appropriate range of possibilities we can expect<br />

to get useful estimates of the average behavior. As mentioned above, some models<br />

employ separate continuity equations for mass and number, called double-moment<br />

schemes, which gives them added flexibility in matching observations, but there are<br />

numerical issues associated with advecting these separately that we would like to<br />

avoid at this time, and so we follow papers such as HDC04 and treat number density<br />

as a diagnostic variable, called the single-moment approach.<br />

We begin with ice crystals, which make up the bulk of clouds in our Jupiter<br />

model. Single-moment bulk microphysics models usually parameterize the number<br />

18


concentration of ice particles, N ice , in terms of local thermodynamic variables. Many<br />

of the existing schemes calculate it from cloud temperature, using the formula given<br />

by Fletcher (1962)<br />

N ice = 0.01e 0.6(T −T0) , (2.5)<br />

where T is the temperature and T 0 is the freezing point. In the <strong>EPIC</strong> model we use<br />

the triple-point temperature, T tp , in equations that call for a freezing point. For the<br />

case of water on Earth, HDC04 note that (2.5) has no physical basis for temperatures<br />

lower than −35 ◦ C and that it often overpredicts the number of small ice crystals in<br />

the upper troposphere. This parameterization increases the number concentration by<br />

an order of magnitude for about every 4 ◦ C temperature drop, so a common practice<br />

is to put an upper bound on the calculated value, e.g. 10 9 m −3 is used in the MM5<br />

model (Grell et al. 1994). On Earth, mid-latitude cirrus clouds often have cloud-top<br />

temperatures colder than −50 ◦ C, and there is a similar problem on Jupiter, where<br />

ammonia clouds typically have temperatures below the ammonia freezing point by<br />

about 50–70 ◦ C. In addition, Galileo Orbiter observations (Gierasch et al. 2000) show<br />

that the strongest convective water clouds on Jupiter can rise above the ammonia<br />

cloud level where the temperature drops below the water freezing point by about<br />

150 ◦ C. Thus, we cannot use (2.5) for either water or ammonia, because it would<br />

produce too many ice crystals in these clouds, causing a bias in the microphysical<br />

functions that use number concentration as a parameter. There are alternatives to<br />

(2.5) that formulate N ice in terms of other variables in addition to temperature, such<br />

as relative humidity or supersaturation (e.g. Rotstayn et al. 2000), but because many<br />

of the coefficients used in these methods are fits to Earth observational data, and the<br />

correlations are not always consistent from case to case (Ryan 2000), we found them<br />

hard to adapt.<br />

19


Instead, to determine N ice we follow HDC04, who obtained reasonable results<br />

by diagnosing ice crystal number concentration from ice mass. Although they also<br />

used several coefficients that are fixed for application to Earth, we find that their<br />

approach is easier to adapt because the parameters are more closely tied to physical<br />

processes. HDC04 proceed by combining two power-law equations for the terminal<br />

velocity, one in terms of diameter and one in terms of mass, a power-law equation for<br />

mass in terms of diameter, and the definition of the mean mass<br />

V t,ice = xD y ice , (2.6)<br />

V t,ice = a(ρ dry q ice ) b , (2.7)<br />

m ice = αD β ice , (2.8)<br />

m ice = ρ dryq ice<br />

N ice<br />

, (2.9)<br />

where y, b, and β are nondimensional exponents, x [m 1−y s −1 ], a [kg −b m 1−3b s −1 ] and<br />

α [kg m −β ] are the corresponding multipliers, and D ice [m] is the diameter of an ice<br />

crystal. To complete the set of equations, we need the number concentration of ice<br />

crystals. In the diagnostic equation the terminal velocity, V t,ice , is equated in (2.6)<br />

and (2.7), and then (2.8) and (2.9) are used to get<br />

N ice = c (ρ dry q ice ) d , (2.10)<br />

where c [kg −d m 3(d−1) ] and d are given by<br />

c = 1 α<br />

( x<br />

a<br />

) β/y<br />

, d = 1 − βb<br />

y . (2.11)<br />

An advantage of this formulation of the single-moment approach is that the microphysical<br />

processes and the advection that affect ice mass will likewise affect the<br />

number concentration (HDC04). HDC04 list values of the coefficients used in the<br />

20


above equations for application to Earth; our task is to estimate appropriate values<br />

for application to Jupiter.<br />

Equation (2.9) does not contain empirical parameters, but equations (2.6),<br />

(2.7), and (2.8) contain a total of 6 parameters that must be fit to observations.<br />

To estimate x and y in (2.6), we apply to cloud-ice crystals the same technique we<br />

use for estimating the terminal fall speed of snow, which involves the correlation of<br />

Reynolds and Davies nondimensional numbers to observations and is described in<br />

Section 2.5.2 below. Equation (2.8) relates mass to diameter and the coefficients can<br />

be calculated based on crystal geometry and species ice-density data. The geometry<br />

a crystal favors is called its habit, and HDC04 tested the effect of crystal habit on<br />

N ice by comparing several different shapes. They found that the choice of ice-crystal<br />

habit does not result in significant differences in N ice and has a small overall impact<br />

on the microphysical behavior. Therefore, for simplicity we follow HDC04 who used<br />

single-bullet crystals, and we adopt their mass- size correlation for water. For other<br />

species, we calculate the coefficients using (2.31).<br />

To calculate the remaining coefficients in the set of equations, at first it appears<br />

that we can proceed either of two ways. In the first, we make assumptions for the<br />

terminal velocity vs. cloud mass equation.<br />

Estimating a and b in (2.7) is rather<br />

difficult because the correlation is more empirical than (2.6). We are not aware of<br />

any fall-speed vs. cloud mass data taken under Jupiter conditions, thus we adopted<br />

the value of b in (2.7) from Earth observations.<br />

The magnitude parameter, a, is<br />

likely to be significantly larger for Jupiter than Earth, judging from various terminal<br />

velocity estimates. In our preliminary calculations we found for cloud-base conditions<br />

that water-cloud particles typically fell about four times faster on Jupiter than Earth,<br />

and that ammonia-cloud particles fell roughly 25% faster than water-cloud particles.<br />

21


Thus, for Jupiter we increased a by a factor of 4 and 5 over its terrestrial value for<br />

water and ammonia, respectively, and we calculated the remaining coefficients, c and<br />

d, for both species using (2.11). However, during our test runs we found that the<br />

threshold mixing ratio for snow autoconversion (described in detail below, see Eq.<br />

2.58) yields a very low value, resulting in too strong and unrealistic snow formation<br />

even for low-density clouds. We conclude that the terminal velocity as a function of<br />

cloud mass on Jupiter is not linearly proportional to its Earth value as it nearly is<br />

for individual particles, thus the method above cannot be used in the current form<br />

in our parameterization. This is an area where laboratory measurements and further<br />

parameterization sensitivity tests are needed.<br />

The other way we can proceed for determining the remaining coefficients is to<br />

assume that the number concentration of particles in the jovian clouds are similar to<br />

those on Earth, and hence to adopt the value of c and d in (2.10) from HDC04. In<br />

the absence of laboratory or in situ data, the terrestrial values of these coefficients<br />

provide at least a reasonable first estimate, but this issue will need to be re-examined<br />

when new data become available. Since coefficients a and b in (2.7) are not part of any<br />

microphysical functions, we do not need to calculate them when using this method.<br />

This method prevailed in the 1D tests described in the next chapter, values of the<br />

cloud densities, the microphysical transfer rates and their time scales are qualitatively<br />

reasonable. The values of the coefficients we used in our simulations are listed in Table<br />

2.2.<br />

2.4 Precipitation Size Distribution<br />

The next task is to account for the wide distribution of sizes of raindrops and<br />

snowflakes. We assume the number density for raindrops follows the exponential form<br />

22


TABLE 2.2<br />

Power law coefficients used in the <strong>EPIC</strong> microphysics code to calculate N ice for water<br />

and ammonia (2.6–2.11). Values needed for the terminal-velocity of rain and snow in<br />

equation (2.21) and (2.29), respectively, are included. Units are MKS.<br />

Coefficient H 2 O NH 3<br />

α ice 7.06165e-3<br />

3<br />

7.06165e-3·ρNH<br />

ρ H2 O<br />

β ice 2.0 2.0<br />

c ice 5.38e+7 5.38e+7<br />

d ice 0.75 0.75<br />

x ice 735.5 702.46<br />

y ice 0.86273 0.8695<br />

x rain 2615.1 2479.0<br />

y rain 0.74245 0.76217<br />

x snow 54..96 53.065<br />

y snow 0.45254 0.46179<br />

ρ liquid 1000.0 733.0<br />

ρ bulk ice 917.0 786.8<br />

23


given by Marshall and Palmer (1948)<br />

n rain (D) dD = N 0rain e −λ rainD dD , (2.12)<br />

where D is the raindrop diameter, n rain dD is the number of raindrops per volume<br />

with diameters in the range D to D + dD, λ rain is the slope parameter [1/m] and<br />

N 0rain is the intercept parameter. Multiplying (2.12) by the mass of each raindrop,<br />

assuming spherical shape and constant density ρ liq , and integrating over all diameters<br />

yields<br />

λ rain =<br />

(<br />

πρliq N 0rain<br />

ρ dry q rain<br />

) 1/4<br />

. (2.13)<br />

Following Earth models (e.g. RH83, D89, FRR96) we use the value of 8 × 10 6 m −4<br />

for the intercept parameter N 0rain for water, and we currently apply the same value<br />

to other species (for lack of better values).<br />

Although snow is more complicated than rain in terms of shape, Gunn and<br />

Marshall (1958) showed that the same functional form for size distribution (2.12) can<br />

be applied to snow if the diameter, D, is interpreted as the melted-snow spherical<br />

diameter. Houze et al. (1979) went further and showed that for precipitation particles<br />

bigger than 1.5 mm in diameter, frozen or otherwise, the exponential size distribution<br />

is quite accurate when the diameter is defined assuming spherical shape and the<br />

density is taken to be the actual rather than melted value.<br />

It has since become<br />

common practice to not convert to melted diameter when applying (2.12) to frozen<br />

hydrometeors in GCM applications, and we follow this practice and use<br />

n snow (D) dD = N 0snow e −λsnowD dD , (2.14)<br />

with<br />

λ snow =<br />

(<br />

πρsnow N 0snow<br />

ρ dry q snow<br />

) 1/4<br />

. (2.15)<br />

24


Ryan (1996) showed that using a constant for the intercept parameter N 0snow (e.g.<br />

RH83, D89) does not approximate well the strong temperature dependence that has<br />

been observed. An error in the number density will lead to inaccuracies in the calculated<br />

rates of sublimation, deposition, accretion of cloud ice by snow, and sedimentation<br />

velocity. Thus, we follow HDC04 and use a temperature-dependent expression<br />

for N 0snow for water that is based on observations by Houze et al. (1979)<br />

N 0snow = (2 × 10 8 m −4 ) min { 0.01 e −0.12(T−Ttp) , 1 } , (2.16)<br />

and we currently apply (2.16) to other species as well (for lack of better functions).<br />

2.5 Fall Speed of Particles<br />

We turn now from number and size distributions to precipitation terminal velocities.<br />

Values for the parameters x and y in the power-law formulation (2.6) have<br />

been determined for rain and snow for different particle sizes and shapes under terrestrial<br />

conditions (Pruppacher and Klett 1997, Rogers and Yau 1989). To generalize to<br />

planetary conditions in the absence of empirical data, we refer to aerodynamic-drag<br />

calculations that are couched in terms of nondimensional numbers. The aerodynamicdrag<br />

force acting on a free-falling particle is defined in terms of the drag coefficient,<br />

C D , as<br />

F D = 1 2 ρ airV 2<br />

t AC D , (2.17)<br />

where A is the area of the particle projected normal to the flow and ρ air is the density<br />

of the gas through which it is falling.<br />

Once the terminal velocity is reached, this<br />

force is exactly balanced by the gravitational force, F G = mg, where m is the mass of<br />

the particle. For strongly viscous flow, meaning low Reynolds number (low-Re) flow,<br />

Stokes’ law provides an analytical expression for C D and hence V t . However, falling<br />

25


precipitation averaged over the complete size range is not a low-Re phenomenon,<br />

therefore we proceed by adapting the methods developed for raindrops and snowflakes<br />

described by Pruppacher and Klett (1997).<br />

2.5.1 Raindrops<br />

The definition of the Reynolds number, Re, may be rewritten in this context<br />

as an expression for the fall speed<br />

where µ air<br />

V t = 1 D<br />

µ air<br />

ρ air<br />

Re , (2.18)<br />

is the dynamic viscosity and Re is defined using diameter (instead of<br />

radius), hence a calculation of Re yields V t . The range of Re that corresponds to fall<br />

speeds that are fast enough for inertial effects to be important, which are neglected<br />

in Stokes’ flow, but slow enough to avoid breakup of a liquid drop is approximately<br />

0.01 ≤ Re ≤ 300, for which raindrops fall like rigid spheres (Pruppacher and Klett<br />

1997). We assume that this range is sufficient to model the fall speed of raindrops for<br />

our purposes. Drop shape and stability are active areas of research, for which surface<br />

tension effects enter in through the nondimensional Weber number.<br />

A discussion<br />

of the current understanding of this complex problem can be found in Chap. 10<br />

of Pruppacher and Klett (1997). Our model is most sensitive to this issue in our<br />

estimate of the maximum drop size (see below), we therefore tested a set of values in<br />

1-D sensitivity tests. Doubling and halving the value of maximum raindrop diameter<br />

in the terminal velocity calculation resulted in less than 5% difference overall between<br />

the times required for the complete rainout in the nominal case (see Chap. 3 below),<br />

thus we conclude that our Jupiter cloud model is insensitive to this parameter.<br />

Given spherical raindrops, the net buoyancy is πD 3 (ρ liq − ρ air )g/6, and when<br />

this is equated to the drag force in terms of C D (2.17), with V t written in terms of<br />

26


Re using (2.18), one obtains<br />

C D Re 2 = 4D3 ρ air g(ρ liq − ρ air )<br />

3µ 2 air<br />

. (2.19)<br />

The new nondimensional number X ≡ C D Re 2 that arises in (2.19) is known as the<br />

Davies number or the Best number. It is significant in the theory of falling precipitation<br />

because, unlike C D or Re, it can be easily evaluated without reference to the<br />

quantity in question, V t , thereby avoiding iteration, while on the other hand, it is a<br />

nondimensional expression of the drag force and hence must have a functional correspondence<br />

with Re. Therefore, the strategy is to find an empirical formula for Re<br />

in terms of X, which will then directly yield V t via (2.18). Clift et al. (1978) determined<br />

empirical correlations for a wide range of Re and X, in which the mass of the<br />

raindrops was converted into diameter assuming rigid sphere geometry. Introducing<br />

W ≡ log 10 X, they found<br />

log 10 Re ≈ −1.81391 + 1.34671W − 0.12427W 2 + 0.006344W 3 , (2.20)<br />

for the interval 12.2 < Re ≤ 6.35 × 10 3 . This range of validity is broad enough to<br />

cover all the raindrop sizes and <strong>atmospheric</strong> conditions we expect for Jupiter. Panel<br />

A) of Fig. 2.1 shows the Reynolds number – Best number correlation.<br />

The next step is to cast V t into the power-law form as a function of raindrop<br />

diameter for Jupiter conditions, V t = xD y . To do this, we calculate the fall speed<br />

of drops in the diameter range 200µm to 5mm (Re of 3.268 to 5260.0 and 1.785 to<br />

4449.0 for water and ammonia, respectively, at the reference pressure level, see below),<br />

and solve for the best-fit x and y (Table 2.2). The lower limit is a typical value<br />

for raindrops, and we do not consider larger drops than the upper limit because they<br />

are likely to break up (Rogers and Yau 1989, Pruppacher and Klett 1997). Such calculations<br />

for Earth usually include a factor that accounts for the change of fallspeed<br />

27


with altitude, for example Foote and DuToit (1969) included the factor (p 0 /p) γ from<br />

empirical correlations, where p 0 = 1000 mbar and γ = 0.4. Their research was limited<br />

to raindrops, but most models use these values to describe the variation of terminal<br />

velocity with height for solid precipitation as well (e.g. RH83, D89). In our calculations<br />

we found the exponent is different for liquid and solid particles and we denote<br />

them with γ liq and γ ice , respectively. On Jupiter, water drops form near the 5 bar<br />

pressure level and ammonia drops are not likely to form at all, but for consistency we<br />

keep p 0 = 1000 mbar and calculate reference theoretical fall speeds at this level. Our<br />

results for Jupiter show γ liq = 0.33 provides a consistently good fit to the altitude<br />

dependence for both water and ammonia raindrops, thus we write the fall speed as<br />

V t = x rain D y rain<br />

(p 0 /p) γ liq<br />

, (2.21)<br />

with γ liq = 0.33 for Jupiter. In panel A) of Fig. 2.2 we compare our results for the<br />

terminal velocity of water raindrops on Jupiter to previously published results for<br />

water raindrops on Earth and methane raindrops on Titan. As expected, the speeds<br />

are larger on Jupiter, which implies that processes such as accretion will be relatively<br />

stronger. Also, the implied rapid vertical transfer of water and other soluble gases<br />

downward are expected to deplete the local abundance of these substances above<br />

Jupiter’s rain evaporation level (Rossow 1978).<br />

The final step is to calculate an average fall speed for rain in an <strong>EPIC</strong>-model<br />

gridbox, which ties this section into the preceding discussion on size distribution. We<br />

follow Srivastava (1967) and express the mass-weighted mean terminal velocity as<br />

V t =<br />

∫ ∞<br />

∫0<br />

∞<br />

0<br />

V t mn dD<br />

, (2.22)<br />

mn dD<br />

where V t (D) is the terminal fall speed, m(D) = πρ liq D 3 /6 is the mass of a raindrop of<br />

diameter D, and ndD is the number density in the diameter range D to D + dD, as<br />

28


Figure 2.1. A) Correlation between the Reynolds- and Best-numbers in the Bestnumber<br />

regime of 10 3 to 10 5 . Solid line represents the function for rain calculated<br />

using (2.20), dashed line shows the correlation for snow calculated using (2.26). B)<br />

Power law curve fit of H 2 O snow terminal velocities at the reference pressure level.<br />

The correlation coefficient of the curve is R = 0.99561.<br />

discussed above. Substituting (2.21) and the distribution function (2.12) into (2.22)<br />

yields the formula we use in the model<br />

V t,rain = (p 0 /p) γ<br />

x rain<br />

(λ rain ) y rain<br />

Γ(y rain + 4)<br />

6<br />

, (2.23)<br />

where Γ(n) = ∫ ∞<br />

0 e −t t n−1 dt is the Gamma function.<br />

2.5.2 Terminal Velocity of Snow<br />

Most of the precipitation in our Jupiter model turns out to be in the form of<br />

snow. As we have already mentioned, snow can be treated in a similar fashion to rain<br />

with slightly less accurate but still satisfactory results. We follow Böhm (1989), who<br />

developed a method for predicting terminal velocities of solid hydrometeors based on<br />

boundary-layer theory with an error ≤ 10% for a wide variety of particles, including<br />

various planar and columnar crystals, rimed and unrimed aggregates, graupel, and<br />

hail. Böhm’s approach readily adapts to planetary situations because of its explicit<br />

29


Figure 2.2. A) Comparison of terminal velocities of raindrops on Earth, Jupiter,<br />

and Titan. The dotted and dashed curves show the calculated fall speed of methane<br />

drops on Titan at 40km altitude and water drops on Earth at the surface, respectively<br />

(Lorenz 1993). The solid curve shows the fall speed of water raindrops on Jupiter at<br />

p = 5.5 bar calculated in this work. B) Comparison of terminal velocities for ammonia<br />

and water snow. The dashed line is H 2 O snow at the surface of Earth from the model<br />

by Dudhia (1989), the dotted line is NH 3 snow at p = 700 mbar from this work, and<br />

the solid line is H 2 O snow on Jupiter at p = 4500 mbar from this work.<br />

30


use of environmental parameters such as density, viscosity, and gravitational acceleration<br />

and of particle parameters such as mass and crystal geometry via a unified<br />

Davies number.<br />

Locatelli and Hobbs (1974) collected data on the mass-size relationship<br />

of solid precipitation particles and we use their correlations in our terminal<br />

velocity calculation. We assume snowflakes to be graupellike snow of hexagonal type,<br />

with density ρ snow , such that the mass may be expressed as<br />

m = 0.333 ρ snow D 2.4 . (2.24)<br />

Locatelli and Hobbs (1974), however, do not list density values for this habit. For<br />

other geometries they found a large variation of density ranging from 20 kg m −3 to<br />

450 kg m −3 . We have no information on density values under Jupiter-like conditions<br />

and for other species, but to take into account the porous nature of snow particles in<br />

this study we use ρ snow = 0.5 ρ bulk ice . We chose this value to take into account for the<br />

graupellike nature of this habit, but future sensitivity tests need to be carried out to<br />

investigate the importance of this parameter.<br />

Böhm (1989) points out that the drag coefficient is much more sensitive to the<br />

holes and irregularities near the edge of an object than to its circumscribed shape,<br />

and that the former can be represented by an effective projected area, A e . He handles<br />

the variety of solid hydrometeor shapes by defining a unified Davies number in terms<br />

of a particle’s mass, effective projected area, and projected area A, which is the area<br />

of the smallest circle or ellipse that completely contains A e ,<br />

X = 8mgρ air<br />

πµ 2 air<br />

( A<br />

A e<br />

) 1/4<br />

. (2.25)<br />

The value of A/A e depends on the crystal type and we have no direct information<br />

on that for planetary applications. From our assumption of hexagonal-shaped snow<br />

particles (A/A e ) 1/4 ≈ 1.05. Böhm inverts X(Re) from his drag-coefficient curve fit to<br />

31


solid precipitation particles to obtain<br />

Re = 8.5<br />

[ (1<br />

+ 0.1519X<br />

1/2 ) 1/2<br />

− 1<br />

] 2<br />

. (2.26)<br />

Panel A) of Fig. 2.1 shows the correlation between the Reynolds number and the<br />

Best number. The fall speed for snowflakes is then obtained as in (2.18), but with<br />

the length scale now generalized in terms of the projected area, A<br />

V t =<br />

( ) π 1/2<br />

µ air<br />

Re , (2.27)<br />

4A ρ air<br />

where Re is defined in terms of the characteristic length. Given the spherical diameter<br />

D, this is simply<br />

V t (D) = 1 D<br />

µ air<br />

ρ air<br />

Re . (2.28)<br />

The next step is to express this V t for snowflakes in a power-law format, as<br />

was done for raindrops above. Based on values found in Pruppacher and Klett (1997)<br />

we calculate the fall speeds of snowflakes in the size range D = .5 to 5 mm (500 to<br />

5000 µm). The lower value is our threshold value for autoconversion of ice crystals<br />

into snow, and with the upper value we take into account the breakup of large solid<br />

precipitation particles. This procedure is done at various pressure levels and then fit<br />

to<br />

V t (D) = x snow D ysnow (p 0 /p) γ ice<br />

, (2.29)<br />

to yield x snow and y snow as given in Table 2.2, with γ ice = 0.47 for both species<br />

and for both ice and snow. As an example, the accuracy of the curve fitting water<br />

snow terminal velocity data is shown in panel B) of Fig. 2.1. Panel B) of Fig. 2.2<br />

compares the terminal velocity of water snow on Earth and Jupiter, and ammonia<br />

snow on Jupiter. As expected, snow falls faster on Jupiter. Interestingly, although<br />

water snowflakes are heavier, they fall much slower than ammonia snowflakes because<br />

the air they encounter is significantly denser.<br />

32


The final step is to calculate the mass-weighted average fall speed for snow in<br />

a gridbox (2.22). One obtains the same form (2.23) as for rain<br />

V t,snow =<br />

x snow Γ(y snow + 4)<br />

(p 0 /p) γ ice<br />

, (2.30)<br />

(λ snow ) ysnow 6<br />

but with the physical differences expressed through the different x, y and λ parameters.<br />

Turning back to the clouds themselves, it is important to distinguish between<br />

the cloud-ice crystals and the much larger snow particles. Following HDC04, in Section<br />

2.4 we derived an equation for the cloud-ice number concentration that requires<br />

the terminal velocity of ice crystals in power-law format. We proceed in the same<br />

manner as for snow, but we find it is best to treat them separately and to fit a different<br />

power law for ice crystals. Following HDC04 we use single-bullet shaped crystals<br />

that have density of the bulk ice, and their value of (A/A e ) 1/4 ≈ 1. We consider<br />

diameters in the range D = 1 to 500µm, and use the diameter-mass relation from<br />

HDC04, which is D = 11.9 m 1/2 , in MKS units. For species other than water, we<br />

scale this by the density such that<br />

D ≈ 11.9<br />

(<br />

ρwater ice<br />

ρ ice<br />

) 1/2<br />

m . (2.31)<br />

The resulting power-law coefficients for ice crystals are included in Table 2.2.<br />

2.6 Microphysical Processes<br />

We next consider the processes that govern phase changes. These affect the<br />

model <strong>dynamics</strong> via the mixing-ratio source/sink terms, Ŝs,p, in the continuity equations<br />

(2.4) and latent heating/cooling terms in the energy equation. Figure 2.3 provides<br />

a graphical overview of the interplay between the phases for a given species like<br />

33


water or ammonia. There are a total of eleven processes that we presently consider.<br />

Dropping the species index, the explicit source terms for phase index p = vap, liq,<br />

ice, rain, and snow are:<br />

Ŝ vap = −P COND − P INT − P DEPI − P REVP − P SEVP , (2.32)<br />

Ŝ liq = P COND + P SMLTI − P RAUT − P RACW , (2.33)<br />

Ŝ ice = P INT + P DEPI − P SMLTI − P SACI − P SAUT , (2.34)<br />

Ŝ rain = P REVP + P RAUT + P RACW + P SMLTS , (2.35)<br />

Ŝ snow = P SEVP + P SACI + P SAUT − P SMLTS . (2.36)<br />

These terms all have units s −1 and the processes corresponding to the abbreviated<br />

subscripts are explained in the caption to Fig. (2.3). The modification to the heating<br />

rate, ˙Q, is<br />

˙Q lat = L evap (P COND + P REVP )<br />

+ L sub (P INT + P DEPI + P SEVP )<br />

+ L fus (P SMLTI + P SMLTS ) , (2.37)<br />

where L evap , L sub , and L fus<br />

are the latent heats of evaporation, sublimation, and<br />

fusion.<br />

2.6.1 Condensation of Vapor/Evaporation of Cloud Liquid (P COND )<br />

The process labeled P COND on the upper-left of Fig. 2.3 represents the condensation<br />

of vapor to form cloud liquid when the vapor mixing ratio exceeds its saturation<br />

value with respect to liquid and the temperature is above the triple point, T ≥ T tp .<br />

Conversely, when the ambient air is subsaturated and T ≥ T tp , then evaporation of<br />

the cloud-liquid phase into vapor takes place. Some cloud microphysical processes<br />

34


Figure 2.3. Schematic diagram of the eleven phase-change processes included in the<br />

five-phase <strong>EPIC</strong> microphysics model. These are P COND : condensation of vapor to<br />

cloud liquid and evaporation of liquid cloud, P INIT : initiation of ice crystals, P DEPI :<br />

deposition of vapor onto ice crystals and sublimation of cloud ice to vapor, P SMLTI :<br />

melting of cloud ice to form cloud liquid, P RACW : collection of cloud liquid droplets by<br />

rain, P RAUT : autoconversion of cloud liquid droplets to rain, P REVP : evaporation of<br />

rain, P SEVP : sublimation and depositional growth of snow, P SMLTS : melting of snow to<br />

form rain, P SACI : collection of cloud ice crystals by snow, and P SAUT : autoconversion<br />

of cloud ice crystals to snow.<br />

35


can be modeled as instantaneous, while others are more accurately treated with a<br />

finite-rate approach. FRR96 represented condensation as a finite-rate process using<br />

P COND = (q satl − q vap )/τ, where q satl is the saturation mixing ratio over liquid and<br />

τ = 100s is the relaxation time for condensation. Since the time constants used in<br />

Earth <strong>modeling</strong> are not necessarily match those on other planets, we found it more<br />

suitable to adapt the formula used in RH83:<br />

⎧<br />

⎨q satl − q vap<br />

P COND = − min<br />

⎩ ∆t<br />

(<br />

) ⎫<br />

−1<br />

1 + L2 vapq satl<br />

, q ⎬<br />

liq<br />

c p R vap T 2 ∆t ⎭ , (2.38)<br />

where L vap is the latent heat of evaporation, c p is the specific heat of air at constant<br />

pressure, R vap is the ideal gas constant of vapor, and ∆t is the timestep used in the<br />

numerical integration. Evaporation is limited by the term q liq /∆t to prevent negative<br />

overshooting of q liq . The latent heat associated with this process in either direction<br />

modifies the saturation mixing ratio, which is taken into account by the correction<br />

factor in parentheses (Teixeira and Miranda 1997).<br />

2.6.2 Initiation of Cloud Ice (P INIT )<br />

The next process is the nucleation of cloud-ice particles from vapor, P INIT .<br />

This is a complicated process and there are many questions that arise in the gasgiant<br />

context that may ultimately require in situ probes to provide the necessary<br />

data for high accuracy. Yair et al. (1995) point out that a key unknown is the nature<br />

of condensation nuclei on Jupiter, and they develop a complete parameterization<br />

to address the problem. At this point, we have opted to use a somewhat simpler<br />

approach for the initialization of cloud ice that fits within our framework, but we<br />

recognize that the importance of this topic will require us and other groups to revisit<br />

the question periodically as new information becomes available.<br />

36


The process of cloud-ice initiation in this study follows that described by<br />

HDC04.<br />

It is a one-way process in which cloud ice is initiated from vapor when<br />

T < T tp , the air is supersaturated with respect to ice (q vap − q sati > 0) and q ice = 0,<br />

where q sati is the saturation mixing ratio over ice. We calculate the rate of initiation<br />

as<br />

{ qice0<br />

P INIT = min<br />

∆t , q }<br />

vap − q sati<br />

, (2.39)<br />

∆t<br />

where q ice0 is the initiated mixing ratio value of cloud ice, and the second term prevents<br />

shooting into ice subsaturation in one time step. Recall that in the algorithm we have<br />

adopted, q ice may be related to the number concentration of ice particles, N ice , via<br />

the powerlaw (2.10) with coefficients c and d. Thus, we may find q ice0 by calculating<br />

the corresponding number concentration N ice0 (which is not to be confused with the<br />

precipitation number-intercept parameters N 0rain or N 0snow used in (2.12), (2.14),<br />

(2.16)), such that<br />

q ice0 = 1<br />

ρ dry<br />

min<br />

{ (Nice0<br />

c<br />

) 1/d<br />

, 1.5 × 10 −3 kg m −3 }<br />

. (2.40)<br />

The upper limit reins in the power law when subject to Jupiter’s temperature range,<br />

which is broader than Earth’s. HDC04 make a point of considering a different formulation<br />

for N ice0 than for N ice in general. For N ice0 , they use Fletcher’s (1962)<br />

temperature-dependent equation to insure that colder temperatures produce more<br />

ice nuclei, but with different coefficients that are specifically chosen to model the<br />

cloud-ice initiation process itself.<br />

We found that the formula given by HDC04 is<br />

more appropriate for our application than Fletcher’s original coefficients (2.5). In<br />

particular, Fletcher’s coefficients produce negligible initiation unless temperatures<br />

are colder than about 30 ◦ C below the freezing point, which on Jupiter translates into<br />

weak water-ice cloud formation below the 2.5 bar level. In contrast, Galileo spacecraft<br />

37


Figure 2.4. Comparison of initialized ice mass calculated from (2.5) for water on<br />

Earth (solid line), and from (2.40) for ammonia (dash-dot line) and water (dotted<br />

line) on Jupiter.<br />

data reveal deep clouds at p ≥ 4 bar (Banfield et al. 1998) and the correlation given<br />

by HDC04 affords such clouds. Yair et al (1995) suggested that the number of ice<br />

nuclei should be on the order of 10 5 m −3 , and to take into account for their suggestion<br />

we have increased the multiplier in HDC04’s equation and we write<br />

N ice0 = max { 10000 e −0.1(T −Ttp) , 1.0 × 10 8 m −3} . (2.41)<br />

In our sensitivity tests (Chaps. 3 and 4 below) we found that varying the value of<br />

the multiplier does not affect the overall results in long time simulations. The upper<br />

limit is for safety such that N ice0 would never grow too large at colder temperatures.<br />

The resulting initialized cloud ice, q ice0 , calculated from (2.40) and (2.41) is shown in<br />

Figure 2.4 .<br />

2.6.3 Deposition of Vapor/Sublimation of Cloud Ice (P DEPI )<br />

Cloud-ice crystals grow by deposition, P DEPI , when the air is supersaturated<br />

with respect to ice, and shrink by sublimation when the air is subsaturated. Given a<br />

38


value of supersaturation, S ice − 1, where S ice = q vap /q sati is the saturation ratio with<br />

respect to ice (also the WMO’s definition of relative humidity), Pruppacher and Klett<br />

(1997) give an expression for the rate of growth of mass by vapor deposition (their<br />

13-76) for a hexagonal-plate ice crystal (their 13-77), which we write in a form similar<br />

to RH83 as<br />

dm<br />

dt ≈ 4D ice(S ice − 1)<br />

A ice + B ice<br />

, (2.42)<br />

where D ice the diameter of the ice crystal, and A ice and B ice are thermodynamic terms<br />

given by<br />

and<br />

A ice = L sub<br />

K air T<br />

B ice =<br />

(<br />

Lsub<br />

R vap T − 1 )<br />

, (2.43)<br />

1<br />

ρ dry q sati χ . (2.44)<br />

Here, K air is the thermal conductivity of air, R vap is the gas constant of vapor, and χ<br />

is the diffusivity of vapor in air. For water vapor in Earth’s atmosphere, RH83 take<br />

χ = 2.26 × 10 −5 m 2 s −1 . However, Pruppacher and Klett (1997) show that assuming<br />

that the transport coefficients K air and χ are constants overestimates the growth rate<br />

of submicron particles.<br />

They cite a size-dependent modification that corrects the<br />

problem (see their 13-14), but then they downgrade the importance of this correction<br />

under natural conditions because it vanishes once the particles grow large enough,<br />

so at the present time we do not include this modification. On the other hand, Yair<br />

et al. (1995) find that the differences between the transport coefficients for Earth<br />

and Jupiter are important to the cloud microphysics, hence we are careful to use<br />

appropriate values. For K air on Jupiter, we use the function given by Hansen (1979).<br />

For χ(T, p), we use the Fuller-Schettler-Giddings correlation with parameters from<br />

Perry’s Chemical Engineers’ Handbook (1997), Tables 5-14 and 5-16, for which the<br />

typical error is less than 5%.<br />

39


Unlike for raindrops and snowflakes, we do not have an expression for the<br />

size distribution of cloud particles, especially not for Jupiter. A common approach<br />

used in Earth models is to assume a monodispersed, average cloud-particle diameter<br />

in each gridbox, D ice , for use in (2.42) and calculate the depositional growth by<br />

P DEPI = N ice dm/dt. We adopt this strategy and calculate D ice using (2.8) and (2.9),<br />

which yields<br />

P DEPI = 4 ¯D i (S ice − 1)N ice<br />

A ice + B ice<br />

= 4 (S ice − 1) α −1/β ice<br />

ice N 1−1/β ice<br />

ice (ρ dry q ice ) 1/β ice<br />

. (2.45)<br />

A ice + B ice<br />

Similarly to the ice initiation process (2.39), we follow the standard practice of limiting<br />

P DEPI such that the deposition rate does not lead to subsaturation in one time step.<br />

In subsaturated conditions, S ice < 1, the right hand side of (2.45) becomes negative<br />

and represents the cloud-ice sublimation rate.<br />

2.6.4 Melting of Cloud Ice/Freezing of Cloud Liquid and<br />

Melting of Snow/Freezing of Rain (P SMLTI )<br />

Next, consider the two processes in Fig. 2.3 that are represented by two-way,<br />

horizontal arrows. As mentioned above, the memory is laid out in the model to handle<br />

mixed-phase situations, but we have decided to wait until the second generation of our<br />

microphysics code before exploring this more complicated option. We currently follow<br />

FRR96 and assume that these melting and freezing processes occur instantaneously<br />

such that no mixed-phase situations arise, in other words that the precipitation in a<br />

gridbox is either rain or snow but not both simultaneously, and the same for cloud<br />

liquid versus cloud ice. Here, cloud ice is completely converted into cloud liquid when<br />

T ≥ T tp and cloud liquid is completely converted into cloud ice when T < T tp<br />

⎧<br />

⎨ q ice /∆T , T ≥ T tp ;<br />

P SMLTI =<br />

⎩<br />

−q liq /∆T , T < T tp ,<br />

(2.46)<br />

40


with a similar conversion between snow and rain<br />

⎧<br />

⎨ q snow /∆T , T ≥ T tp ;<br />

P SMLTS =<br />

⎩<br />

−q rain /∆T , T < T tp .<br />

(2.47)<br />

2.6.5 Autoconversion of Cloud Liquid to Rain (P RAUT )<br />

One of the two processes by which cloud liquid can be converted into rain is<br />

through mutual coalescence, also called autoconversion and labeled P RAUT in Fig. 2.3.<br />

During this process, cloud droplets collide with each other as a result of their nearly<br />

random motion. Similarly to RH83 and FRR96, we have implemented the approach<br />

introduced by Kessler (1969)<br />

P RAUT = max { α RAUT (q liq − q liq0 ), 0 } , (2.48)<br />

where α RAUT is a rate coefficient and q liq0 is a threshold value for autoconversion. Weinstein<br />

(1970) found that this parameterization is sensitive to q liq0 , but fortunately<br />

for planetary applications, is not so sensitive to α RAUT . Since neither parameter is<br />

constrained by observations for planetary atmospheres, we have adopted the terrestrial<br />

values used by FRR96, q liq0 = 0.00025 kg kg −1 and α RAUT = 0.001 s −1 . These<br />

parameters may need to be reexamined before applying them to a planet where rain<br />

is an important component of the hydrological cycle (such as Titan).<br />

2.6.6 Collection of Cloud Liquid by Rain (P RACW )<br />

Cloud droplets are also swept up by falling raindrops, which is the process<br />

labeled P RACW in Fig. 2.3. Similarly to RH83 and FRR96, we calculate the mass<br />

increase of a single raindrop of diameter D due to collection of cloud liquid as<br />

dm<br />

dt = π 4 D2 V t (D) q liq E rl , (2.49)<br />

41


where E rl is the cloud-liquid collection efficiency for rain, which we take to be unity<br />

as is usually done. We substitute the fallspeed of the raindrop V t (D) from (2.21),<br />

multiply the equation with the Marshall-Palmer distribution (2.12), and integrate<br />

over all diameters to get an expression similar to that in D89<br />

P RACW = π ( ) γliq<br />

4 x p0 Γ (y rain + 3)<br />

rainq liq E rl N 0rain<br />

p λ y . (2.50)<br />

rain+3<br />

rain<br />

2.6.7 Deposition of Vapor on Rain/Evaporation of Rain (P REVP )<br />

In addition to the two processes just discussed, raindrops also grow by deposition<br />

of vapor when they pass through a supersaturated environment, and conversely<br />

shrink by evaporation when they fall into a subsaturated environment. This is the<br />

process labeled P REVP in Fig. 2.3. The growth rate of a single raindrop may be expressed<br />

with an equation analogous to (2.42) for the deposition of vapor on cloud<br />

ice, P DEPI (see Byers 1965; Pruppacher and Klett 1997, p. 547), but for the spherical<br />

shape of the raindrop, and with a ventilation factor F that accounts for the drop’s<br />

motion<br />

dm<br />

dt = 2πDF (S liq − 1)<br />

A rain + B rain<br />

, (2.51)<br />

where S liq = q vap /q satl is the saturation ratio over liquid. Analogous to (2.43) and<br />

(2.44), we have<br />

and<br />

A rain = L vap<br />

K air T<br />

B rain =<br />

(<br />

Lvap<br />

R vap T − 1 )<br />

, (2.52)<br />

1<br />

ρ dry q satl χ . (2.53)<br />

We use the ventilation factor given by Beard and Pruppacher (1971)<br />

F = f 1r + f 2r Sc 1/3 Re 1/2 , (2.54)<br />

42


where Sc = (µ air /ρ dry )/χ is the Schimdt number, µ air is the dynamic viscosity, and<br />

the empirically determined parameters are f 1r = 0.78 and f 2r = 0.308. Integrating<br />

(2.51) over all drop sizes using the Marshall-Palmer distribution, with Re in terms of<br />

the diameter power-law formulation for the terminal velocity (2.21)<br />

Re = ρ air<br />

DV t = ρ ( ) γliq<br />

air p0<br />

x rain D 1+y rain<br />

, (2.55)<br />

µ air µ air p<br />

yields an expression similar to that in D89<br />

P REVP = 2π(S [ ( ) 1/2 (<br />

liq − 1)N 0rain f1r<br />

+ f<br />

A rain + B rain λ 2 2r Sc 1/3 xrain ρ air p0<br />

rain<br />

µ air p<br />

) γliq /2 ]<br />

Γ(c r )<br />

, (2.56)<br />

λ cr<br />

rain<br />

where c r = (y rain + 5)/2. In case of subsaturation, S liq < 1, the right hand side of<br />

(2.56) is negative and is the evaporation rate.<br />

2.6.8 Autoconversion of Cloud Ice to Snow (P SAUT )<br />

The final three processes in Fig. 2.3 involve snow and have forms similar to the<br />

analogous processes involving rain. First, the autoconversion of cloud ice to snow,<br />

P SAUT , occurs when the mixing ratio for cloud ice exceeds a threshold q ice0<br />

P SAUT = max { α SAUT (q ice − q ice0 ), 0 } . (2.57)<br />

Even though the threshold concept is the same, having two liquid drops merge into<br />

one is quite a different process from having two solid crystals stick together, and so it<br />

is common to treat the autoconversion rate coefficients for rain and snow differently.<br />

For snow, RH83 and HDC04 simply use α SAUT = 1/∆t such that it is treated as an<br />

instantaneous process, and we currently do the same. Since we do not yet know the<br />

maximum size of cloud-ice particles in planetary settings, we follow these papers and<br />

assume D ice0 = 500 µm, above which autoconversion to snow occurs. Substituting<br />

43


this value into (2.8), and combining it with (2.9) and (2.10) yields the threshold<br />

q ice0 = 1<br />

ρ dry<br />

(<br />

cice α ice D β ice<br />

ice0<br />

) 1/(1−dice )<br />

. (2.58)<br />

2.6.9 Collection of Cloud Ice by Snow (P SACI )<br />

Falling snow collects cloud ice just as falling rain collects cloud liquid, and the<br />

form of the mass equation is the same<br />

dm<br />

dt = π 4 D2 V t (D)q ice E si , (2.59)<br />

where E si is the cloud-ice collection efficiency for snow. Since the accretion of cloud<br />

ice by snow is a stronger process at higher temperatures (Lin et al. 1983), we have<br />

adopted the temperature-dependent ice-collection efficiency function used by HDC04<br />

E si = e 0.05(T −Ttp) . (2.60)<br />

Given the terminal velocity (2.29) and number distribution (2.14), integrating over<br />

all particle sizes generates an equation analogous to (2.50) for rain<br />

P SACI = π ( ) γice<br />

4 x p0 Γ (y snow + 3)<br />

snowq ice E si N 0snow<br />

p λsnow<br />

ysnow+3 . (2.61)<br />

2.6.10 Deposition of Vapor on Snow/Sublimation of Snow (P SEVP )<br />

The final microphysical process we consider is the interaction between snow<br />

and vapor, P SEVP . This is similar to the interaction between cloud ice and vapor,<br />

except that snow experiences a ventilation effect because of its motion and has an<br />

exponential size distribution. As is standard, we model the ventilation effect using<br />

(2.54) with f 1s = 0.65 and f 2s = 0.44 (Thorpe and Mason 1966). After accounting for<br />

the size distribution, the process equation resembles (2.56) for rain with a geometry<br />

44


factor of 4 as in (2.42) instead of 2π<br />

P SEVP = 4(S [ ( ) 1/2 (<br />

ice − 1)N 0snow f1s<br />

+f<br />

A snow + B snow λ 2 2s Sc 1/3 xsnow ρ air p0<br />

snow<br />

µ air p<br />

) γice /2 ]<br />

Γ(c s )<br />

, (2.62)<br />

λ cs<br />

snow<br />

where c s = (y snow + 5)/2. In case of subsaturation, S ice < 1, the right hand side of<br />

(2.62) is negative and it represents the sublimation rate.<br />

2.7 Code Implementation<br />

We end this chapter with some details concerning the implementation of the<br />

microphysics scheme in the <strong>EPIC</strong> model. The following chapters detail our initial<br />

results from application of this model to Jupiter using one- and two-dimensional<br />

configurations.<br />

2.7.1 Positivity adjustment<br />

It is important for physical and computational reasons that all mass-related<br />

variables, including all five phases of each active species, remain positive definite everywhere<br />

(Ooyama 2001). For this reason, we use a positive-definite advection scheme<br />

in the dynamical core and a robust conditional-branching function for floating point<br />

numbers in the microphysics code (Belding, 2000). Truncation errors throughout the<br />

model may still lead to the occasional incidence of negative mass, and so we take the<br />

standard precaution of performing an adjustment step to make the masses positive<br />

definite. There are various ways to do this, for example borrowing from nearby gridboxes,<br />

but after trying various schemes in the end we have decided to simply reset<br />

negative values to zero. The main drawback of this simple approach is a lack of strict<br />

mass conservation, but we note that the algorithms we use are designed to conserve<br />

total energy and enstrophy on the grid rather than total mass. In practice, for any<br />

45


easonable resolution the total mass in the model does not drift in an appreciable<br />

manner and we have not encountered any problems with this issue.<br />

2.7.2 Sequence of Process Transfers<br />

Implementing cloud microphysics in a GCM is an inherently complicated task,<br />

and in addition to establishing good algorithms for each process, the order in which<br />

these processes are applied to the model variables (the timesplitting) usually affects<br />

the outcome, primarily because more than one process competes for the same material.<br />

There is little agreement yet among research groups as to what order microphysical<br />

processes should applied. For example, FRR96 apply freezing and melting<br />

processes first and Morrison et al. (2003) apply them last. Therefore, no model description<br />

is complete without a description of its process sequencing, which we include<br />

here. Similarly to FRR96, we calculate the freezing and melting processes first, which<br />

we treat as instantaneous, and we update the mixing ratios accordingly, but do not<br />

update the temperature at this point. As mentioned above, in our current implementation<br />

“warm” and “cold” processes do not take place simultaneously in a given<br />

grid box. For the case T > T tp , we calculate P COND first, followed by P RAUT , P RACW<br />

and P REVAP , in this order. For the case T < T tp , we calculate P INT or P DEPI first,<br />

depending on whether cloud ice has previously existed or not at the grid location,<br />

and then P SAUT , P SACI and P SEVP . We update the mixing ratios and the temperature<br />

via the heat equation after all these transfer rates are calculated. As has been described<br />

above, care is taken to avoid unnatural flip-flopping between supersaturation<br />

and subsaturation from one timestep to the next, and to avoid negative mass. After<br />

the microphysical processes have been applied, in the dynamical core we advect each<br />

phase of each active species, including the relative fall speeds of snow and rain. The<br />

46


next chapter describes tests and results of the model when configured for the vertical<br />

dimension only.<br />

47


CHAPTER 3<br />

1D SIMULATIONS<br />

3.1 Introduction<br />

Before using the microphysics scheme in a full GCM model it is useful to<br />

test its functions in simplified simulations, where we examine the robustness of the<br />

processes, modify or fine-tune the parameterizations, and evaluate the importance of<br />

individual elements of the scheme. The Earth <strong>atmospheric</strong> <strong>modeling</strong> community uses<br />

single column models for this purpose (Randall and Cripe, 1999). We have tested the<br />

microphysics scheme in a simplified, one-dimensional version of <strong>EPIC</strong>, the setup and<br />

results of which are described in this chapter.<br />

We have tested our cloud scheme in two steps. First, we have established a<br />

nominal model and investigated its runtime properties. In the absence of <strong>dynamics</strong><br />

we perturb the model with a high initial supersaturation that generates a thermodynamically<br />

unbalanced state, which triggers the various processes into action. The<br />

description of the column model and the results from these simulations are discussed<br />

in Section 3.2 and 3.3, respectively. Second, we have explored the sensitivity of the<br />

model using a suite of initial abundance and relative humidity values. These details<br />

are given in Section 3.3.1. Section 3.3.2 describes the effect of clouds on the static<br />

stability of the atmosphere.<br />

48


3.2 Nominal <strong>Model</strong> Description<br />

The one-dimensional model is built with 155 layers, which is much higher<br />

vertical resolution than we can typically employ in full three-dimensional simulations.<br />

The base of the model is located at the 10 bar pressure level and the top of the model<br />

is at 10 mbar. When vertical layers were spaced proportional to the logarithm of<br />

pressure, which corresponds to approximately even spacing in height, or when they<br />

were evenly spaced in pressure, the vertical resolution was not fine enough within the<br />

saturation range of at least one of the species, thus we have opted for a custom spacing<br />

that provides sufficient number of layers in the expected cloud and precipitation<br />

region and a coarser resolution at the top and bottom of the model for computational<br />

efficiency. Figure 3.3 shows the position of the layers. The top five layers of the<br />

model are used as a sponge to damp the reflection of vertically propagating waves.<br />

In the hybrid vertical coordinate system (Dowling et al. 2006) a σ-to-θ coordinate<br />

transition pressure of p σ<br />

= 120 mbar (for the undisturbed model) was chosen for<br />

all the 1D simulations (Fig. 3.1). For this vertical resolution we use a time step of<br />

∆t = 30 s for all of the simulation in this chapter. This value was determined by<br />

decreasing the time step until it was the largest possible that did not introduce any<br />

numerical instability.<br />

3.2.1 Vertical Structure of <strong>Model</strong> Environment<br />

The initial temperature profile above the visible cloud deck at about the 700<br />

mbar level was taken from Voyager radio occultation data published by Lindal et al.<br />

(1981). Below this pressure level our temperature profile followed a dry adiabatic<br />

extrapolation of the Voyager data. Figure 3.1 shows the mean temperature profile<br />

49


Figure 3.1. (Left) The solid line represents the basic-state temperature profile used<br />

in our simulations. Data were given by Lindal et al. (1981) with a dry adaiabtic<br />

extrapolation towards the bottom. The dotted line shows the variation of potential<br />

temperature in the model, the dashed line is the hybrid vertical coordinate, ζ, with a<br />

transition to a pressure based coordinate below p σ = 120 mbar. (Right) Mean zonal<br />

wind profile of the basic-state, as described in the text.<br />

50


used in our calculations.<br />

The only in situ measurement of the zonal wind profile below the cloud level<br />

comes from the Galileo probe doppler-wind measurement data. The probe entered<br />

one of Jupiter’s singular regions, a so-called 5µm hot spot (Orton et al. 1998), a<br />

phenomenon in Jupiter’s equatorial region that is thought to be caused by the dry<br />

downdraft of an equatorially trapped wave (Ortiz et al. 1998, Showman and Dowling<br />

2000). Such a downdraft modifies the local thermal and wind profiles, thus the data<br />

should probably not be generalized for the whole planet.<br />

In our test simulations,<br />

we found that the model is sensitive to the deep wind profile, causing numerical<br />

instability for cases of too much vertical shear in regions of low static stability. In<br />

this study we have reduced the vertical wind shear to 20% of the value measured by<br />

the Galileo probe. Above the cloud deck the magnitude of the zonal wind is set to<br />

follow the thermal-wind decay determined by Gierasch et al. (1986) from Voyager<br />

IRIS data. The resulting zonal wind, u, as a function of pressure can be written as:<br />

⎧<br />

0, p ≤ p 0 · e −cs ,<br />

(<br />

⎪⎨ u(p<br />

u(p) = 0 ) 1 − 1 ln p )<br />

0<br />

, p 0 · e −cs < p < p 0 ,<br />

c s p<br />

( [ ] )<br />

u(p)probe<br />

⎪⎩ u(p 0 ) du vert − 1 + 1 , p 0 ≤ p,<br />

u(p 0 )<br />

(3.1)<br />

where p 0 = 680 mbar, c s = 2.4 is the mean value of the scale height of the wind shear<br />

for the planet (Gierasch et al. 1986), u(p) probe is the wind profile measured by the<br />

Galieo probe (Atkinson et al. 1997), and du vert = 0.2. This profile used is shown in<br />

Figure 3.1.<br />

3.2.2 Vertical Structure of Clouds<br />

The first experiments to simulate the vertical structure of Jupiter’s atmosphere<br />

used one-dimensional thermochemical equilibrium models (Lewis 1969, Wei-<br />

51


denschilling and Lewis 1973). These models assumed solar-based compositions. It is<br />

widely used and convenient to compare the abundance of elements to a solar value,<br />

which represents the ratio of the number of the species atoms to hydrogen atoms in<br />

the Sun. In this study we have adopted the standard solar values published by Anders<br />

and Gravesse (1989). The Galileo probe in situ measurements and ground based observations<br />

both question the validity of the solar composition of the planet. A recent<br />

review of Jupiter’s composition is given by Taylor et al. (2004). The analyses of the<br />

probe radio signal attenuation data suggests a deep ammonia abundance of 4.0 ± 0.5<br />

times the solar abundance (Folkner et al. 1998). Updated Galileo probe mass spectrometer<br />

(GPMS) measurement analysis provides a 2.96 ± 1.13 solar enrichment for<br />

the 8.9-11.7 bar pressure region (Wong et al. 2004). The probe could not identify<br />

the deep value of water abundance, and based on the dry nature of hot spots, the<br />

measured sub-solar value serves only as a lower limit.<br />

The condensible volatiles we include in this study are ammonia and water.<br />

Although Rossow (1978) has argued that ammonia and water will form a single condensing<br />

vapor solution and its saturation vapor pressure can be a more complicated<br />

function of temperature, following Stoker (1986) and Yair et al. (1995) we have represented<br />

water as a pure substance in these simulations. Sulfur is also important and<br />

is thought to produce an NH 4 SH cloud that lies between the NH 3 cloud above and<br />

the H 2 O cloud below. However, the chemistry involved represents an added level of<br />

complexity, and we have decided to delay the treatment of sulfur to a future paper.<br />

The volume molar mixing ratios generated by the <strong>EPIC</strong> model for solar abundance<br />

are 0.000187 and 0.00142 for NH 3 and H 2 O, respectively. In the nominal model we<br />

used a 4.0 times solar deep ammonia abundance and, in the absence of indisputable<br />

evidence, a solar deep water abundance. In the <strong>EPIC</strong> model, species are initialized<br />

52


Figure 3.2. <strong>Model</strong>s of the vertical profiles of ammonia (top) and water (bottom) volume<br />

mixing ratios (v.m.r.’s) used in this study. The solid lines indicate deep v.m.r.’s<br />

of 4.0 bounded by the saturation curve and the cold trap (see text). The v.m.r.’s<br />

corresponding to solar values of 0.5 and 1.0 are indicated with finely dotted and<br />

dashed-triple dotted lines, respectively. Saturation curves of 120% (supersaturated)<br />

and 80% (subsaturated) are represented by the coarsely dotted and dashed parallel<br />

curves, respectively.<br />

53


starting in the abyssal layer and working up. Based on the temperature of the layer, a<br />

saturation mole fraction value is calculated and limited by the user-defined maximum<br />

allowed saturation factor (e.g. 1.1 in our nominal 1D case). The final mole fraction of<br />

the species in the layer is the minimum of this value and the deep abundance value.<br />

Proceeding upwards, once we pass the tropopause the temperature begins to increase,<br />

and in this case, instead of increasing the abundance according to the temperature,<br />

we use the value of the previous layer (the “cold trap” model). Figure 3.2 shows the<br />

resulting initial vapor profile.<br />

3.3 Results of 1D Simulations<br />

The nominal 1D model was used to test the formation and evolution of clouds<br />

and precipitation as governed by the microphysical processes described in the previous<br />

chapter. As expected, the cloud initialization starts at the first time step as a direct<br />

result of our 10% supersaturation. Ammonia-ice cloud forms at and above the 820<br />

mb pressure level, while water-ice cloud forms at and above the 4700 mb pressure<br />

level. Interestingly, a single layer of liquid-water cloud forms at the 4750 mb pressure<br />

level, which has a slightly higher temperature than the triple point of water. These<br />

results show good agreement with published thermochemical models. Atreya el al.<br />

(1999) predicted an ammonia ice cloud with a base at about 720 mb and a waterammonia<br />

solution cloud at 5690 mb assuming solar abundance for each species. The<br />

difference between their predicted cloud base levels and our results comes from the<br />

use of slightly different initial T(p) profiles. The cloud initialization does not remove<br />

all the supersaturation, thus the remaining excess vapor provides a source for fast<br />

and strong crystal growth that can be observed at the lower water-ice cloud region.<br />

Parallel with the growth of the clouds, snow formation starts near the ammonia-ice<br />

54


cloud base and in the midst of the water-ice cloud, and rain forms in the liquid-water<br />

cloud. At about 10 minutes into the model run, snow starts to form near the waterice<br />

cloud base as well, and then melts into rain as it falls through the triple-point<br />

temperature level.<br />

As the falling precipitation reaches the lower cloud levels, the<br />

increasing amount of snow collects more and more ice particles, thus we have the<br />

highest value of snow density at about one hour into the simulation. This occurs near<br />

the water-ice cloud base with a peak value of 3.8 × 10 −4 kg m −3 . The extremum of<br />

water-ice cloud density is 2.57 × 10 −4 kg m −3 and liquid-water cloud has a maximum<br />

density of 6.27×10 −4 kg m −3 . The maximum value of water rain density is 9.7×10 −5<br />

kg m −3 , and it reaches down to about the 5200 mb pressure level before it completely<br />

evaporates. The maximum density for the ammonia-ice cloud and snow is 4.3 × 10 −4<br />

kg m −3 and 8.83×10 −5 kg m −3 , respectively, and the precipitation falls down to about<br />

the 880 mb pressure level before it evaporates. The fallout of ammonia precipitation<br />

is complete at about 100 minutes into the simulation, while all the water rain and<br />

snow fall out and evaporate in about 5 and 8 hours, respectively.<br />

The maximum temperature increase in a single layer due to latent heating is<br />

0.073 K in the ammonia-cloud region and 0.192 K in the water-cloud region. The<br />

cooling caused by the evaporating precipitation is -0.0196 K and -0.458 K below the<br />

ammonia and water cloud base, respectively. For ammonia, the absolute value of the<br />

cooling is smaller than the heating because only a small portion of the cloud turns<br />

into evaporating precipitation. This is in agreement with Carlson et al. (1988), who<br />

suggested that ammonia clouds on Jupiter are non-precipitating or weakly precipitating.<br />

On the other hand, for water the temperature change of cooling in a single layer<br />

is higher than for the heating. Figure 3.3 shows that this is because the water clouds<br />

and precipitation form over a widespread region between about 3.5 and 5 bars (∼14<br />

55


Figure 3.3. The time evolution of water and ammonia clouds and precipitation in<br />

the nominal 1D model of Jupiter’s atmosphere. Blue represents the water-ice cloud,<br />

green the liquid-water cloud (too thin to be seen here, but clearly visible on Fig. 3.4c),<br />

and tan the ammonia-ice cloud. Yellow would indicate a liquid-ammonia cloud, but as<br />

expected, this phase does not form in the model. Snow is marked with black outlines,<br />

and red outlines are rain. The initial abundance of ammonia and water vapor is 1.0<br />

and 4.0 times solar, respectively, with 10% initial supersaturation. The lowest values<br />

plotted for both the filled contours and the outlines correspond to mass mixing ratios<br />

of 10 −6 kg m −3 and are incremented by 0.33 kg m −3 . Black ticks on the right hand<br />

side indicate model layer-interface positions.<br />

56


km in height) such that the amount of heat generated by condensation is distributed<br />

vertically, whereas the evaporation of precipitation from all of those layers mostly<br />

takes place in a relatively narrow region at about 5 bars.<br />

<strong>EPIC</strong> is a hydrostatic model such that the vertical velocity is a diagnostic<br />

variable. Scaling the hybrid vertical velocity, ˙ζ, into m/s shows that the maximum<br />

updraft and downdraft velocities (gridbox average) caused by the latent heat are 1.77<br />

mm s −1 and -0.94 mm s −1 , occurring at the 3770 mb and at the 5020 mb pressure<br />

level, respectively.<br />

3.3.1 Sensitivity to abundance<br />

While the dryness observed by the Galileo probe can be explained by the<br />

unique nature of the 5 µm hot spots, other observations also contradict the idea that<br />

the abundance of condensible species in the cloudy region is similar to the deep abundances,<br />

in fact there appear to be a range of abundance values at different altitudes.<br />

From Voyager radio occultation measurements, Lindal et al. (1981) reported a volume<br />

mixing ratio of 2.21 × 10 −4 at the 1 bar pressure level, which is slightly higher<br />

than the solar value. From Voyager IRIS spectra, Carlson et al. (1993) suggested<br />

an ammonia vapor profile that decreases with height and is sub-saturated above the<br />

condensation level, based on a fit to data at all wavelengths. Analysis of Jupiter’s microwave<br />

emission indicates depleted ammonia on a global scale, with ammonia mixing<br />

ratios decreasing with altitude at p < 4 bar, and with about 0.5 times solar ammonia<br />

at p ≤ 2 bar (de Pater et al. 2001). There are fewer observational constraints on the<br />

water clouds, since they are below the visible cloud deck and there is a wide range<br />

of suggested abundance values, varying from subsolar to ten times solar. For more<br />

complete listings of various observational data see Sromovsky et al. (1998) and Taylor<br />

57


et al. (2004).<br />

We have carried out experiments to explore the effect of variation of the deep<br />

vapor abundance and the initial supersaturation on the cloud structure. Results for<br />

0.5, 1.0, and 4.0 times solar abundance for both ammonia and water are summarized<br />

here (see Fig.3.2). Figure 3.4 shows the cloud evolution for these three deep abundance<br />

values, each initialized with 10% supersaturation. As expected, the location of the<br />

cloud bases is controlled by this parameter. Water-snow sublimates instead of first<br />

melting into rain when we reduce the water abundance to 0.5 times solar.<br />

The<br />

presence of supercooled liquid is probably required for the lightning activity on Jupiter<br />

observed by Voyager 1 (Smith et al. 1979) and by Galileo (Little et al. 1999),<br />

in analogy with Earth. Based on our tests we estimate that at the latitudes that<br />

exhibit lightning on Jupiter a minimum value of approximately solar local abundance<br />

is needed. The 4.0 times solar abundance results in a thick liquid cloud under the<br />

water-ice cloud (Fig. 3.4c). Large raindrops can form in this dense precipitation field,<br />

which collects liquid cloud drops very effectively. Figure 3.4c shows that the accretion<br />

process is so strong that the liquid portion of the water cloud almost vanishes on the<br />

order of 1 hour, thus we expect any such dense clouds on Jupiter to be local and very<br />

short lived. Ammonia clouds do not precipitate in our simulations when the deep<br />

abundance is only solar or subsolar.<br />

Figure 3.5 illustrates the effect of varying the initial supersaturation, with the<br />

deep mixing ratios held to the nominal-model values. This parameter controls how<br />

dense a cloud becomes. Below 3% initial supersaturation level we found that neither<br />

species precipitates. At 3%, rain forms in the liquid (warm) region of the water cloud<br />

and it falls out completely in about half an hour. In the ice cloud, at first crystals<br />

grow by collecting vapor particles, until around 50 minutes into the simulation they<br />

58


Figure 3.4. Cloud and precipitation evolution using various initial deep abundance<br />

values, with initial supersaturation set to 10%. Panel A) has a deep abundance value<br />

of 0.5 solar for both ammonia and water, panels B) and C) have deep abundances for<br />

both species of 1.0 solar and 4.0 solar, respectively. Color coding is as in Fig. 3.3.<br />

The lowest values plotted for both the filled contours and the outlines correspond to<br />

mass mixing ratios of 10 −6 kg m −3 and are incremented by 0.5 kg m −3 .<br />

59


Figure 3.5. Sensitivity of cloud and precipitation evolution to various initial supersaturation<br />

values, with the deep abundance values set to the nominal case of 4.0 and<br />

1.0 times solar for ammonia and water, respectively. Panel A) has a supersaturation<br />

level of 3% for both ammonia and water, panels B) and C) have supersaturation levels<br />

of 6% and 12%, respectively. Color coding is as in Fig. 3.3. Contours are as in Fig.<br />

3.4.<br />

60


grow big enough to form snow. Increasing the initial supersaturation reduces this<br />

time, and for sufficiently high initial values precipitation forms instantly (Fig. 3.5b<br />

and c). The time required for complete fallout of rain and snow also increases with<br />

the supersaturation.<br />

The ammonia cloud requires a high supersaturation level to<br />

form precipitation, consistent with expectations.<br />

3.3.2 Static Stability of the Atmosphere<br />

The stratification of Jupiter’s atmosphere plays a significant role in controlling<br />

its <strong>dynamics</strong>.<br />

Stratification can be expressed in terms of the buoyancy or Brunt-<br />

Väisäla frequency, N, or its square. Because of the generalized definition of potential<br />

temperature used in the <strong>EPIC</strong> model, the Brunt-Väisäla frequency does not satisfy the<br />

classic meteorological relation N 2 = (g/θ)dθ/dz, and so we use the formula derived<br />

in the appendix of Dowling et al. (2006)<br />

N 2 = g 2 [<br />

−ρβ + α2 T<br />

c p<br />

+ dρ ]<br />

env<br />

, (3.2)<br />

dp<br />

where where α ≡ −ρ −1 (∂ρ/∂T ) p = ρ(∂v/∂T ) p is the isobaric volume thermal expansivity,<br />

v ≡ 1/ρ is the specific volume, and β ≡ ρ −1 (∂ρ/∂p) T<br />

is the isothermal<br />

compressibility. When N 2 is positive the atmosphere is statically stable, when it is<br />

zero the atmosphere is neutral and the temperature profile is dry adiabatic, while<br />

negative N 2 is the characteristic of an unstable atmosphere.<br />

We have turned on and off various components of the microphysics scheme in<br />

order to test their contribution to the effects of the active moisture in the atmosphere.<br />

Table 3.1 summarizes which phases and their microphysical processes were used in<br />

the simulations. Figure 3.6 shows the comparison between an initial state with no<br />

condensibles (i.e.<br />

dry model, Case 1) and the nominal model prior to any latent<br />

61


TABLE 3.1<br />

Microphysical components turned on and off in the Brunt-Väisälä test runs (see text<br />

for details).<br />

Vapor Clouds Precipitation Latent heat<br />

Case 1 - - - -<br />

Case 2 + - - -<br />

Case 3 + + + -<br />

Case 4 + + - +<br />

Case 5 + + + +<br />

Figure 3.6. Brunt-Väisälä frequency of the initial state. The solid line represents<br />

an initial state with no moisture added (Case 1 in Table 3.1) and the dashed line<br />

corresponds to the initial state of the nominal model (Case 2), for which the deep<br />

volume mixing ratios are 4.0 and 1.0 times solar for ammonia and water, respectively.<br />

The dotted line shows the variation of the molar mass of moist air with height in the<br />

nominal model.<br />

62


heating or precipitation effect (Case 2). Below each cloud base, at about 800 mb<br />

and 5000 mb for ammonia and water, respectively, the contribution of the added<br />

mass to the stratification is negligible, but above in each cloud condensation region<br />

there is a significant increase in N 2 from the increased molar mass. We found a<br />

similar but smaller stabilizing effect in the low abundance cases, and as previously<br />

mentioned the cloud bases move somewhat higher in the atmosphere. Figure 3.7<br />

compares the static stability of the atmosphere at 2 hours into the nominal model<br />

simulation. In the case of non-precipitating clouds (Case 4) the effect of the latent heat<br />

of condensation increases the stability in the region where clouds form, but the change<br />

is less significant than from just the molar mass effect.<br />

The case of precipitation<br />

without including latent heating (Case 3) moves the stable region downwards as the<br />

snow and rain falls, and introduces a slight instability in the deeper zone where the<br />

precipitation evaporates. Using the full microphysics (Case 5) has the same effect as<br />

the precipitation-only case but the magnitudes of both the stability and instability<br />

are increased by the added effect of latent heating. In the ammonia cloud region,<br />

the sensitivity of the static stability to turning on and off any of the components of<br />

the microphyisics is found to be negligible compared to the molar mass effect. Based<br />

on these tests we confirm that water effects Jupiter’s <strong>atmospheric</strong> <strong>dynamics</strong> not only<br />

through the latent heating, but also significantly from the redistribution of mass by<br />

precipitation. Next, we examine how the cloud microphysics interacts with Jupiter’s<br />

belts and zones in a two-dimensional, meridional-plane version of the model.<br />

63


Figure 3.7. Brunt-Väisälä frequency at 2 hours into the simulation with selected<br />

microphysics components turned off. The dotted line represents a model with the<br />

microphysics turned off (Case 2 in Table 3.1), moisture plays the role of a tracer<br />

with mass that is otherwise passive in this simulation. The dashed line shows the<br />

effect of the latent heating turned off in the model (Case 3), while the dot-dashed<br />

line indicates the effect of no precipitation (Case 4). The solid line represents the full<br />

<strong>EPIC</strong> microphysics scheme (Case 5).<br />

64


CHAPTER 4<br />

MERIDIONAL PLANE SIMULATIONS<br />

4.1 Introduction<br />

In terrestrial meteorology, moist thermo<strong>dynamics</strong> leads to fascinating but complicated<br />

phenomena, from simple precipitation to supercell thunderstorms to hurricanes<br />

to global wave patterns like the Madden-Julien oscillation. The rich set of observations<br />

we have for Jupiter’s atmosphere imply that jovian moist thermo<strong>dynamics</strong><br />

is equally as interesting, and equally as complicated. With the cloud-microphysics<br />

scheme we have developed and tested in 1D, it is now possible to follow many new<br />

lines of inquiry, for example the interactions of jovian anticyclones and cyclones with<br />

moisture, and the effects of moist thermodyamics on the acceleration of jet streams.<br />

However, our present goal is to test the processes in the new cloud microphysics package<br />

in as simple and controlled a manner as possible, thus we leave for future work<br />

the full-scale investigations just mentioned.<br />

In this chapter we set up experiments in a two-dimensional model to observe<br />

the latitudinal variation of the cloud structure. Having a full 2D meridional plane<br />

opens up the possibility of many interesting flow patterns that cannot be realized<br />

with the 1D model. We describe cases here that illustrate the typical behavior of the<br />

model, with the emphasis placed on stressing the algorithms rather than on achieving<br />

a completely natural state. In our simulations we use the term“storm” to distinguish<br />

65


on the one hand from the ubiquitous, long-lived jovian vortices, and on the other hand<br />

from localized convective thunderstorms that have a shorter time scale and would<br />

require a nonhydrostatic mesoscale or regional-scale model to properly simulate.<br />

4.2 <strong>Model</strong> Description<br />

In our two-dimensional Jupiter model we use 45 layers. The base of the model<br />

is located at the 8 bar pressure level and the top of the model is at 10 mbar. Similarly<br />

to our 1D simulations, the layers are custom-spaced to give higher resolution in the<br />

cloudy regions. Figure 4.2 shows the position of the layers. We use the top four layers<br />

as sponge layers, and p σ = 120 mbar as the σ-to-θ coordinate transition pressure for<br />

the undisturbed model.<br />

The nominal model is a channel spanning 40 ◦ S to 40 ◦ N<br />

latitude, with 161 evenly spaced grid points (0.5 ◦ resolution). To avoid numerical<br />

instability a time step of ∆t = 120 s is used for all of the 2D simulations in this<br />

chapter. The hyperviscosity coefficient used to damp high-frequency computational<br />

modes is ν 8 = (0.01)[(1/2400)∆y 8 /∆t] m 8 s, and the value of divergence damping is<br />

set to ν div = (0.5)[(1/30)∆y 2 /∆t] m 2 s .<br />

4.2.1 Vertical Structure of <strong>Model</strong> Environment<br />

Our wind profile is similar to the one described by (3.1) except that it is<br />

expanded by a latitudinal dependence to yield a two dimensional wind field:<br />

⎧<br />

0, p ≤ p 0 · e −cs ,<br />

(<br />

⎪⎨ u(λ, p<br />

u(λ, p) =<br />

0 ) 1 − 1 ln p )<br />

0<br />

, p 0 · e −cs < p < p 0 ,<br />

c s p<br />

( [ ] )<br />

u(p)probe<br />

⎪⎩ u(λ, p 0 ) du vert − 1 + 1 , p 0 ≤ p,<br />

u(p 0 )<br />

(4.1)<br />

where λ is the planetographic latitude, and similarly to the 1D experiments, p 0 = 680<br />

mbar, c s = 2.4, and du vert = 0.2. The velocity profile of the jets as a function of<br />

66


Figure 4.1. (Top) Zonal wind profile of the initial state. Solid and dashed contours<br />

represent eastward and westward jets, respectively. The plotted contours are u =<br />

[−50, −30, −10, 10, 30, 50, 70, 100, 130] m s −1 , with the positive values corresponding<br />

to eastward winds. (Bottom) Mean zonal temperature profile of the basic state, units<br />

are K.<br />

67


latitude at the p 0 pressure level is the standard Limaye (1986) profile from Voyager<br />

cloud-tracking observations. We kept the unknown vertical wind shear at the 20%<br />

value used in Chapter 3, thus du vert = 0.2.<br />

In our 2D simulations we start with the same vertical temperature profile that<br />

was used in the 1D case (Section 3.2). This sounding profile is applied at the equator,<br />

and gradient wind balance dictates the thermal profile in the rest of the model. The<br />

resulting zonal temperature and wind profiles are shown in Fig. 4.2<br />

4.2.2 Vertical Structure of Condensibles<br />

Similarly to the 1D simulations, we have established a nominal abundance<br />

model and through sensitivity tests have investigated how different values of the<br />

species abundances and initial saturation ratios affect the structure of clouds. To<br />

be consistent with our 1D case, in the 2D nominal model the ammonia and water<br />

abundances are set to 4.0 times and 1.0 times solar, respectively.<br />

4.3 Results<br />

First, we investigated the cloud formation and evolution in the meridional<br />

plane. In the first 200 days of our simulations we ran the model with the microphysics<br />

turned off. The mass of the water and ammonia vapor fields is present and<br />

gets advected with the winds, but otherwise these condensibles are passive. In this<br />

geostrophic-adjustment phase the temperature and meridional velocity values differ<br />

significantly from the quasi steady-state values, and since most of the microphysics<br />

functions are a strong function of temperature (e.g. saturation vapor pressure), if we<br />

allowed cloud formation during this adjustment period, the resulting cloud structure<br />

would contain unwanted spurious clouds that are produced on a non-physical basis.<br />

68


Figure 4.2. Vertical structure of clouds in the nominal 2D model at 1 day into the<br />

active microphysics simulation. Color coding and contours are as in Fig. 3.3. Black<br />

ticks on the right hand side indicate layer-interface positions.<br />

At the end of this 200 days period, the model reaches a sufficiently balanced state<br />

that we can turn the cloud microphysics on without triggering unrealistic structure.<br />

Although it is an option in the <strong>EPIC</strong> model, in these simulations we choose not to<br />

re-initialize the vapor field at the end of this adjustment phase. During the sensitivity<br />

experiments we have tested the re-initialized vapor field case too, and the results are<br />

similar, except that a few days time delay is experienced.<br />

In the nominal model, we use a 10% initial supersaturation level that ensures<br />

rapid cloud growth and precipitation formation. A snapshot of the clouds at day 1 of<br />

69


the active microphysics simulation is shown in Figure 4.2. At this time, the maximum<br />

densities of the ammonia and water ice clouds are 0.044 kg m −3 and 0.076 kg m −3 ,<br />

respectively.<br />

The cloud and precipitation fields for water start out more uniform<br />

than for ammonia, but cloudy and cloud free regions develop for both species. The<br />

meridional structure reflects the vapor redistribution during the first 200 days of<br />

passive advection and geostrophic adjustment. The cloud base of the water-ice clouds<br />

is at about the 4500 mbar pressure level, while the ammonia cloud base is at about<br />

the 800 mbar level, consistent with the 1D simulation results of Chapter 3. Figure 4.3<br />

shows the latitudinal variation of cloud density for water and ammonia, sampled at<br />

4000 mbar and 750 mbar, respectively, for 1 and 30 days into the active simulation.<br />

While the median value of cloud density is about 0.02 g m −3 for both species, the<br />

deviations of the ammonia values are significantly larger, such that cloudy and cloud<br />

free regions alternate frequently.<br />

One conspicuous difference between the 2D and 1D simulations is the location<br />

of the ammonia clouds near the northern edge of the domain. This phenomenon is<br />

caused by the high velocity jet in the vicinity of 24 ◦ N, coupled with our assumptions<br />

about the vertical shear given by (4.1). The gradient balance of the high velocity jet<br />

creates a “shift” in the temperature field at about the ammonia cloud level that can<br />

be seen on the bottom panel of Figure 4.1. The temperature is 5-15 ◦ K higher than at<br />

other latitudes at the same pressure level, which raises the saturation vapor pressure<br />

such that the cloud base gets shifted upwards by about 200 mbar, resulting in nonuniform<br />

bases and tops for the ammonia clouds in the meridional plane. This clearly<br />

illustrates that the zonal-wind structure strongly affects the vertical position and the<br />

overall structure of the clouds. Regarding precipitation, the ammonia snow falls out<br />

in about 10 hrs, while the water snow/rain precipitates out everywhere except in the<br />

70


equatorial region in about 2 days. This is longer than the time we observed in the<br />

1D case. In those simulations the lack of horizontal motion prevented air flows to<br />

feed the cloudy regions with moist air, while in 2D we have seen circulations to form<br />

(for example, two closed circulations on top left panel of Fig. 4.6) that increased the<br />

moisture level in certain regions, causing a longer time for the complete fallout of<br />

precipitation in those areas.<br />

A detailed look at the nominal model reveals an interesting result. Panel A) of<br />

Figure 4.4 shows that at about 6 days into the active simulation a storm starts to form<br />

at the equator, while the rest of the domain stays in a rather steady state No other<br />

dynamic weather events of a similar magnitude were observed in these tests, consistent<br />

with low velocities outside the equatorial region. However, these weak winds do make<br />

the clouds more homogeneous over time throughout the whole domain (Fig. 4.3).<br />

Figure 4.5 shows the temperature change at day 40 into the active simulation<br />

compared to the beginning of the active simulation. In general, where cloud formation<br />

occurred we can see a net positive change, while below the cloud bases, where the precipitation<br />

evaporated, the atmosphere has cooled down as expected. The maximum<br />

value of heating and cooling are 0.66 ◦ K and −0.67 ◦ K, respectively, located near the<br />

base of the rising cloud at about 3000 mbar and below the liquid water layer-clouds of<br />

the equatorial storm at about 5500 mbar, respectively. This equatorial storm system<br />

is one of the more interesting features that has come out of our model testing, and<br />

so we take a closer look at it next.<br />

4.3.1 Equatorial Storm System<br />

Figure 4.6 shows details of the equatorial storm. Near the 5000 mbar level<br />

the meridional winds move air towards the equator in both hemispheres, which re-<br />

71


Figure 4.3. Temporal evolution of the latitudinal variation of cloud density for<br />

ammonia (top) and water (bottom), sampled at the 750 mbar and the 4000 mbar<br />

pressure levels, respectively. Panels on the left show clouds at 1 day into the active<br />

microphysics simulation, panels on the right show the same at 30 days. The dense<br />

region of the water clouds were lifted by the latent heat and buoyancy effects during<br />

this period causing a depletion at the reference pressure level, so we have included<br />

the dotted line on panel D, which represents the new dense layer of water clouds at<br />

the 3500 mbar level.<br />

72


Figure 4.4. Time evolution of the equatorial storm in the nominal model. Color<br />

coding and contours are as in Fig. 3.3.<br />

73


Figure 4.5. 40 day difference of zonal mean temperature compared to the initial<br />

state of the active simulation. Solid and dashed contours represent increases<br />

and decreases in temperature, respectively. The plotted contours are ∆T =<br />

[−0.25, −0.2, −0.1, −0.05, −0.01, 0.01, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3] ◦ K. Color coding<br />

and contours of clouds are as in Fig. 3.3.<br />

74


sults in the formation of solid and liquid stratus clouds between 1 ◦ and 3 ◦ latitude<br />

in both hemispheres.<br />

The equatorial updraft of the forming circulation produces<br />

a large, dense water-ice cloud at the equator.<br />

As this cloud rises in the updraft,<br />

more condensation occurs that strengthens the updraft, which is qualitative confirmation<br />

that our microphysical processes are functioning appropriately. In addition,<br />

the downdrafts of the storm at about ±2 ◦ latitude drag down drier and less dense air<br />

from about the 3000 mbar pressure level, diluting the lower-level water clouds. This<br />

dilution eventually yields cloud free regions at those latitudes (Fig. 4.6). We can<br />

also see a circulation between the 5000 and 7000 mbar pressure levels, rotating in the<br />

opposite direction. The updrafts of this circulation feed the stratus clouds with extra<br />

moisture, creating a dense and precipitating cloud system. At the ammonia cloud<br />

level the dominant motion is horizontal (the vertical component of the wind vectors<br />

is significantly smaller than those in the water cloud region). The ammonia clouds of<br />

the storm system do not precipitate.<br />

The latent heat from the cloud formation and the buoyancy caused by the<br />

continuous precipitation starts to lift the equatorial storm at about 6 days into the<br />

active cloud simulation. The lifting creates more cloud and precipitation formation<br />

that will yield stronger updraft in the cloudy region and stronger downdraft around<br />

it. The maximum and minimum values of the vertical velocity are 0.138 m s −1 and<br />

-0.063 m s −1 , respectively, both occurring at 17 days into the active simulation. The<br />

center of the rising cloud and the equatorward-side of the stratus clouds both form<br />

snow. In case of the equatorial storm, the snow falls almost a scale height before<br />

melting into rain and finally evaporating at about the 5250 mbar pressure level.<br />

The fallout of precipitation and the rising of the cloud ends at about 30 days<br />

into the active simulation. In the resulting cloud system, the cloud top of the equa-<br />

75


Figure 4.6. Close up images of the equatorial storm. Panels on the left show clouds<br />

and wind velocity vectors, panels on the right show clouds and precipitation. Color<br />

coding and contours are as in Fig. 3.3. The vertical component of the velocity vectors<br />

is exaggerated by a factor of 200 to enhance clarity.<br />

76


torial storm is at about the 1000 mbar level while the cloud base is at about 3000<br />

mbar. There is a cloud-free region around and below this storm, spanning about ±2 ◦<br />

latitude. Near the 5000 mbar pressure level we observe some thin stratus clouds; both<br />

solid and liquid phase clouds are present. A dense ammonia cloud system covers this<br />

equatorial storm throughout the simulation, with a cloud top at about 500 mbar and<br />

the cloud base at about 800 mbar. An ammonia-cloud free region spans from about<br />

5 ◦ to 7 ◦ latitude on both hemispheres at this level.<br />

4.3.2 Sensitivity tests<br />

In general, when the microphysics is fully activated in the model, the response<br />

to tests involving unstable initial conditions (supersaturation) takes the form of largescale<br />

storm systems. The location and evolution of these storms varies depending on<br />

the nature and severity of the initial inbalance. Similarly to the 1D case, experiments<br />

are carried out here to investigate how the model responds to various parameter<br />

settings. These tests are of an engineering nature and meant to stress the system,<br />

but they nevertheless reveal the potential of the model’s new functionality.<br />

Figure 4.7 illustrates how varying the initial saturation ratio affects the model<br />

when the deep abundance values are kept the same as in the nominal case. Results<br />

using saturation ratio values of S = 0.95 , 1.00 , and 1.05 are shown. During the<br />

passive simulation (0-200 days), the equatorial <strong>dynamics</strong> redistributed enough vapor<br />

to form some ammonia clouds at the beginning of the active simulation, even in the<br />

case of a subsaturated initial condition. The top row of Figure 4.7 illustrates that in<br />

the S = 0.95 case, only a trace of water cloud forms at the later stage of the simulation.<br />

In the case of an exact 100% saturation initial condition, we can see all the familiar<br />

components of the equatorial storm system known from the nominal simulation, which<br />

77


Figure 4.7. Zonal cloud structure with modified initial saturation ratio values.<br />

Arrows indicate traces of water clouds. Color coding and contours are as in Fig. 3.3.<br />

78


has a 110% initial saturation, including the rising cloud with the heavy precipitation,<br />

the stratus clouds on both sides of the equatorial storm at about the 5000 mbar<br />

pressure level, and the continuous ammonia cloud cover. Besides this storm system<br />

four other cloudy regions formed in this simulation, with strong vertical coherence<br />

between the two cloud species. The approximate locations of these ammonia and<br />

water clouds are between -36 ◦ and -33 ◦ , between -26 ◦ and -20 ◦ , between 18 ◦ and<br />

22 ◦ , and a smaller water cloud region between 33 ◦ and 35 ◦ latitude. Interestingly,<br />

if we compare these locations with the zonal wind profile (Fig. 4.8) we see in this<br />

simulation that these cloudy areas coincide with the zones (anticyclonic wind shear<br />

regions). The S = 1.05 case (bottom row of Fig. 4.7) shows a very similar picture to<br />

our nominal model except near the northern edge of our domain, we see more cloud<br />

free areas in the region where the clouds are elevated. As expected based on our 1D<br />

tests, the cases with reduced initial saturation ratio yield less dense clouds.<br />

In all the simulations the equatorial region is the most active. Since this model<br />

domain is symmetric about the equator we pose an obvious test question, which is<br />

whether this activity is indeed tied to the equator or perhaps just to the middle<br />

of the domain. To answer this, we set up experiments where the latitude range is<br />

not centered on the equator. Figure 4.9 shows a simulation between -20 ◦ and 60 ◦<br />

latitude, with everything else the same as the nominal model. In this case, a similar<br />

equatorial storm system formed to that in the nominal model. This suggests that<br />

Jupiter’s equatorial <strong>dynamics</strong> is distinct from its mid latitude <strong>dynamics</strong> and plays<br />

an important role in shaping the meridional cloud structure, just as is the case for<br />

Earth.<br />

Figure 4.10 shows a reduced deep abundance simulation. In Section 3.3.1 we<br />

discussed that the deep abundance values of 4.0 times solar and 1.0 times solar for<br />

79


Figure 4.8. Cloud and zonal wind structure at 30 days into the active simulation<br />

using S = 1.00 initial saturation ratio. Color coding and contours of clouds and<br />

precipitation are as in Fig. 3.3. Contours of zonal winds are as in Fig. 4.1<br />

Figure 4.9. Sensitivity test in the domain of -20 ◦ to 60 ◦ latitude. Color coding and<br />

contours are as in Fig. 3.3.<br />

80


Figure 4.10. Results with reduced abundance initial condition. Color coding and<br />

contours are as in Fig. 3.3.<br />

ammonia and water, respectively, are close to the observed upper bounds. Thus, to<br />

set up an abundance test we use a reduced profile of 1.0 times solar and 0.5 times solar<br />

for NH 3 and H 2 O, respectively. We keep the initial saturation ratio at S = 1.10, the<br />

same as in the nominal case. Similarly to the 1D experiments, we find that the deep<br />

abundance value controls the location of the cloud base. The location of the ammonia<br />

cloud base has changed from about 800 mbar to about 600-650 mbar and the base<br />

of the water clouds has changed from about 4500 mbar to about 4000 mbar. The<br />

density of the ammonia clouds has reduced and we observe more cloud free regions.<br />

The equatorial storm develops at a later stage of the simulation, with the main lifting<br />

event beginning at about 50 days into the active microphysics simulation, compared<br />

to about 6 days for the nominal case.<br />

In all the simulations where the equatorial storm appears, the rising cloud<br />

stops with its cloud top near the 1000 mbar region, slightly below the ammonia cloud<br />

base level. Is it the ammonia cloud that stops the rise of the water cloud Figure<br />

4.11 illustrates a test case where we include only water, with everything else the<br />

81


same as in the nominal model. The right panel shows that the storm event is very<br />

similar to that in the nominal case. The location and the time scale of the storm<br />

are the same, also the fact that the cloud does not rise above the 1000 mbar region.<br />

In other words, the increase in static stability in the upper troposphere compared<br />

to the lower troposphere is sufficient to significantly suppress vertical motions, with<br />

or without the presence of active ammonia vapor and clouds. Consistent with this<br />

observation, on the close-up images of the nominal case (Fig. 4.6) we see that in<br />

the upper troposphere the vertical component of the wind vectors is small. We can<br />

conclude that the <strong>atmospheric</strong> constituents above the 1000 mbar level will in general<br />

have little interaction with the deeper regions. Observations indicate that in certain<br />

localized regions, water clouds penetrate through the ammonia clouds (Ingersoll et al.<br />

2000), as shown in Figure 1.2. Such events are the result of local convective storms<br />

that create updrafts of 50-100 m s −1 , and presumably this high velocity carries the<br />

water clouds through the stable region of the atmosphere. Our large-scale, hydrostatic<br />

model creates updrafts of the order of 0.1 m s −1 , thus in order to model nonhydrostatic<br />

convective clouds, new mesoscale components will eventually need to be added to the<br />

<strong>EPIC</strong> model.<br />

Finally, as with our 1D experiments, we also test cases where selected components<br />

of the microphysics are turned off. The left panel of Figure 4.12 shows a<br />

model with the latent heating turned off, and the right panel shows a model with<br />

the precipitating phases turned off. The most significant difference from the nominal<br />

model is that the equatorial storm system does not appear in either case. Using a<br />

hot-air balloon analogy, this means that it is not enough just to pull the chain for<br />

fire, or to throw the sandbags out of the basket, we need to do both at the same time<br />

in order for this hot-air balloon to rise. Since one of our goals in this project is to<br />

82


Figure 4.11. Snapshots from the model, where water was the only condensible<br />

species. Color coding and contours are as in Fig. 3.3.<br />

Figure 4.12. Cloud structure in model runs with selected microphysical components<br />

turned off, at 40 days into the active simulation. The left panel shows a model with the<br />

latent heating turned off, while the right panel shows a model with the precipitating<br />

phases turned off. Color coding and contours are as in Fig. 3.3.<br />

83


determine which components of a cloud scheme are needed in jovian simulations, we<br />

conclude that any future Jupiter model that employs only a simple, non-precipitating<br />

cloud scheme, or neglects latent heating, will not be able to reliably simulate even<br />

the large-scale cloud <strong>dynamics</strong> accurately.<br />

84


CHAPTER 5<br />

CONCLUSIONS<br />

An active hydrological cycle has been added to the <strong>EPIC</strong> general circulation<br />

<strong>atmospheric</strong> model (GCM). To date, this research project is the first that has developed<br />

a fully functional cloud microphysics scheme for a global-scale GCM for gas giant<br />

applications, and is particularly significant because it has the capability of <strong>modeling</strong><br />

multiple species at the same time. Recently published Earth water-cloud schemes<br />

have been adapted and altered to determine appropriate planetary values for various<br />

microphysical functions and parameters. The new subroutines calculate saturation<br />

vapor pressure, microphysical transfer rates, latent heating and cooling, and massaveraged<br />

terminal velocities. The resulting computer program can be downloaded as<br />

open-source software from NASA’s Planetary Data System (PDS) Atmospheres node.<br />

To verify that our numerical algorithms perform reasonably we have carried<br />

out a set of experiments with an emphasis on Jupiter. During these simulations our<br />

goal has been to explore the behavior of the model and to ensure that it is properly<br />

matching expectations based on previous studies, such as Jupiter’s cloud positions and<br />

the nature of precipitation on gas giants, while providing leads for the more complete<br />

and more realistic investigations now made possible by this new functionality.<br />

Both in 1D and in 2D we have tested our microphysics scheme in two steps.<br />

First, we have established a nominal case and observed its runtime properties. In<br />

the nominal case simulations we have used 4.0 times solar and 1.0 times solar values<br />

85


for the deep abundance for ammonia and water, respectively. We have perturbed<br />

the models with a large initial supersaturation, and the resulting thermodynamically<br />

unbalanced state has triggered various microphysical processes.<br />

Second, we have<br />

tested the behavior of the scheme by setting up a suite of sensitivity tests. During<br />

these experiments we have varied the initial saturation and abundance values, and<br />

turned on and off selected components of our cloud scheme.<br />

In our 1D single column test cases, the ammonia-ice cloud formed at and<br />

above the 820 mbar pressure level, water-ice cloud formed at and above the 4700<br />

mbar pressure level, and a single layer of liquid-water cloud formed at the 4750 mbar<br />

level. These results are in agreement with published thermochemical models. During<br />

the sensitivity tests we have observed that the value of the deep abundance controls<br />

the cloud base level, while the value of the initial supersaturation affects the density<br />

of the cloud, as expected. We have shown how turning selected components off in the<br />

cloud scheme affects the static stability.<br />

Our 2D meridional plane cases have allowed us to test how different configurations<br />

affect the latitudinal variation of the clouds. The bases of the clouds for both<br />

species matched those in the one-dimensional simulations, except at 24 ◦ N latitude,<br />

where the ammonia cloud base was raised by about 200 mbar as a result of gradient<br />

balancing in the vicinity of Jupiter’s high velocity jet. In general, water clouds tended<br />

to be more uniform than the ammonia clouds, for which cloudy and cloud-free regions<br />

alternated frequently. We have seen only a small amount of motion in most parts<br />

of the domain, but this motion was significant in that it made both layers of clouds<br />

more uniform in density. We have observed an interesting storm system as a result of<br />

the equatorial <strong>dynamics</strong> of the model atmosphere. This heavily precipitating storm<br />

raised the water clouds to about the 3000 mbar level, with the cloud top at about 1000<br />

86


mbar. We have shown that the stability of the upper troposphere tends to stop any<br />

mixing between large air masses below and above this level. This equatorial region<br />

has shown the most activity in all of our sensitivity tests. The time required for the<br />

storm to form was controlled by values of the initial deep abundance and saturation<br />

ratio. Significantly, in the 2D case we have shown that turning major components of<br />

the cloud scheme off results in no equatorial storm.<br />

Important conclusions drawn from these experiments are that the locations of<br />

cloud base and cloud top are sensitive to the underlying zonal-wind structure, and that<br />

the simultaneous inclusion of both latent heat and precipitating phases is necessary<br />

in order to simulate important large-scale weather phenomena. Additional research<br />

opportunities are now made possible by the new microphysics scheme. The model is<br />

ready to be applied for short-term weather simulations where the radiation does not<br />

affect the results significantly, such as analysis of the interaction of jovian vortices<br />

with moisture. The interacting vapor, cloud ice, cloud liquid, rain, and snow phases<br />

are ready to be coupled to new development efforts planned for the <strong>EPIC</strong> model. In<br />

particular, they allow for the development of a realistic radiation transfer scheme, a<br />

subgrid-scale cumulus parameterization, and the inclusion of non-hydrostatic terms.<br />

Each of these will directly affect and improve the cloud simulations.<br />

Addition of<br />

<strong>atmospheric</strong> chemistry is straightforward and will make it possible to simulate the<br />

ammonium-hydrosulfide clouds, as well as tackle problems involving stratospheric<br />

photochemistry.<br />

Future work directly related to this dissertation will likely include the extension<br />

of the microphysics code to other planets, including active methane for Uranus,<br />

Neptune, and Titan, sulphuric acid for Venus, and dust for Mars. The analysis of incoming<br />

data from ongoing planetary missions, such as <strong>atmospheric</strong> measurements of<br />

87


the Cassini-Huygens mission to Saturn and Titan, and the European Space Agency’s<br />

Venus Express mission to Venus, are motivating the continued development of planetary<br />

general circulation <strong>atmospheric</strong> models with realistic hydrological cycles.<br />

There is much that can be done to make improvements to the cloud microphysics<br />

model developed here. Laboratory experiments are needed to clarify the difference<br />

between certain key properties of terrestrial and jovian clouds. For example,<br />

droplet size spectra and collection efficiencies need to be measured in cloud-chamber<br />

tests that reproduce planetary conditions and include species other than just water<br />

(such as ammonia). The incorporation of data from additional Earth-based and<br />

spacecraft observations and in-situ measurements will add more constraints on the<br />

global distribution of condensibles and on the zonal-wind structure. We support a<br />

multi-probe planetary mission to Jupiter that would add valuable information on the<br />

vertical and latitudinal structure of winds, temperatures, and abundances below the<br />

visible cloud deck.<br />

At this point, we do not claim that the results in this dissertation represent<br />

exact simulations of the clouds on Jupiter. However, this research has demonstrated<br />

that a GCM has the potential to model planetary weather patterns in a reasonable<br />

fashion, and it has laid the groundwork to further explore some of the most fascinating<br />

<strong>atmospheric</strong> features in the Solar System. Our work represents an important step<br />

towards a better understanding of these phenomena.<br />

88


REFERENCES<br />

[1] Anders, E. and N. Gravesse, 1989. Abundances of the elements - Meteoritic and<br />

solar. Geochim. Cosmochim. Acta 53, 197–214.<br />

[2] Atkinson, D.H., A.P.Ingersoll, and A. Seiff, 1998. Deep winds on Jupiter as<br />

measured by the Galileo probe. Nature 388, 649–650.<br />

[3] Atreya, S.K., M.H. Wong, T.C. Owen, P.R. Mahaffy, H.B. Niemann, I. de Pater,<br />

and P. Drossart, 1999. A comparison of the atmospheres of Jupiter and Saturn:<br />

deep <strong>atmospheric</strong> composition, cloud structure, vertical mixing, and origin.<br />

Planet. Space Sci. 47, 1243–1262.<br />

[4] Banfield, D., P.J. Gierasch, M. Bell, E. Ustinov, A.P. Ingersoll, A.R. Vasavada,<br />

R.A. West, and M.J.S. Belton, 1998. Jupiter’s cloud structure from Galileo imaging<br />

data. Icarus 135, 230–250.<br />

[5] Beard, K.V., and H.R. Pruppacher, 1971. A wind tunnel investigation of the<br />

rate of evaporation of small water droplets falling at terminal velocity in air. J.<br />

Atmos. Sci. 28, 1455–1464.<br />

[6] Belding, T.C., 2000. fcmp Homepage. http://fcmp.sourceforge.net/ (12 December<br />

2004).<br />

[7] Böhm, H.E., 1989. A general equation for the terminal fall speed of solid hydrometeors.<br />

J.Atmos.Sci. 46, 2419–2427.<br />

89


[8] Bohren, C.F., and B.A. Albrecht, 1998. <strong>Atmospheric</strong> thermo<strong>dynamics</strong>. Oxford<br />

University Press, 402 pp.<br />

[9] Byers, H.R., 1965. Elements of cloud physics. University of Chicago Press, 191<br />

pp.<br />

[10] Carlson, B.E., W.B. Rossow, G.S. Orton, 1988. Cloud microphysics of the giant<br />

planets. J.Atmos.Sci. 45, 2066–2081.<br />

[11] Carlson, B.E., A.A. Lacis, and W.B. Rossow 1993. Tropospheric gas composition<br />

and cloud structure of the Jovian North Equatorial Belt. J. Geophys. Res. 98,<br />

5251-5290.<br />

[12] Clift, R., J.R. Grace, and M.E. Weber, 1978. Bubbles, Drops, and Particles.<br />

Academic Press, 380 pp.<br />

[13] Del Genio, A.D., and K.B. McGrattan, 1990. Moist convection and the vertical<br />

structure and water abundance of Jupiter’s atmosphere. Icarus 84, 29–53.<br />

[14] de Pater, I., D. Dunn, P. Romani, and K. Zahnle, 2001. Reconciling Galileo<br />

probe data and ground-based radio observations of ammonia on Jupiter. Icarus<br />

149, 66–78.<br />

[15] Dowling, T.E., A.S. Fischer, P.J. Gierasch, J. Harrington, R.P. LeBeau, Jr.,<br />

and C.M. Santori, 1998. The Explicit Planetary Isentropic-Coordinate (<strong>EPIC</strong>)<br />

<strong>Atmospheric</strong> <strong>Model</strong>. Icarus 132, 221–238.<br />

[16] Dowling, T.E., E. Colon, J. Kramer, R.P. LeBeau, G. C. Lee, R. Morales-<br />

Juberías, Palotai Cs.J., V.K. Parimi and A.P. Showman, 2006. The <strong>EPIC</strong> at-<br />

90


mospheric model with an isentropic/terrain-following hybrid vertical coordinate.<br />

Icarus 182, 259-273.<br />

[17] Dudhia, J. 1989. Numerical study of convection observed during the winter monsoon<br />

experiment using a mesoscale two-dimensional model. J. Atmos. Sci. 46,<br />

3077–3107.<br />

[18] Ferrier, B.S. 1994. A double-moment multiple-phase four-class bulk ice scheme.<br />

Part I.: Description. J. Atmos.Sci. 51, 249–280.<br />

[19] Fletcher, N.H., 1962. The physics of rainclouds. Cambridge University Press, 386<br />

pp.<br />

[20] Folkner, W.M., R. Woo, and S. Nandi, 1998. Ammonia abundance in Jupiter’s<br />

atmosphere derived from the attenuation of the Galileo probe’s radio signal. J.<br />

Geophys. Res. 103, 22,847–22,855.<br />

[21] Foote, G.B., and P.S. DuToit, 1969. Terminal velocity of raindrops aloft.<br />

J.Appl.Meteor. 8, 249–253.<br />

[22] Forget, F., F. Hourdin, and O. Talagrand, 1998. CO2 snowfall on Mars: simulation<br />

with a general circulation model. Icarus 131, 302–316.<br />

[23] Fowler, L.D., D.A. Randall, and S.A. Rutledge, 1996. Liquid and ice cloud microphysics<br />

in the CSU General Circulation <strong>Model</strong>, Part I: <strong>Model</strong> description and<br />

simulated microphysical processes. J. Climate 9, 489–529.<br />

[24] Garcia-Melendo, E., A. Sánchez-Lavega, and T.E. Dowling, 2005. Jupiter’s 24 ◦<br />

N highest speed jet: Vertical structure deduced from nonlinear simulations of a<br />

large-amplitude natural disturbance. Icarus 176, 272–282.<br />

91


[25] Gierasch, P.J., B.J. Conrath, and J.A. Magalhães, 1986. Zonal mean properties<br />

of Jupiter’s upper troposphere from Voyager infrared observations. Icarus 67,<br />

456–483.<br />

[26] Gierasch, P.J., A.P. Ingersoll, D. Banfield, S.P. Ewald, P. Helfenstein, A. Simon-<br />

Miller, A. Vasavada, H.H.Breneman, D.A. Senska, and the Galileo Imaging<br />

Team, 2000. Observation of moist convection in Jupiter’s atmosphere. Nature<br />

403, 628–630.<br />

[27] Grell, G.A., J. Dudhia, and D. Stauffer, 1994. A description of the fifthgeneration<br />

Penn State/NCAR Mesoscale <strong>Model</strong> (MM5). NCAR Tech. Note<br />

NCAR/TN-398+STR, 117 pp.<br />

[28] Gunn, and J.S.Marshall, 1958. The distribution with size of aggregate snow<br />

flakes. J.Meteor. 15, 452–461.<br />

[29] Hansen, C.F., 1979. Viscosity and thermal conductivity of model Jupiter atmospheres.<br />

NASA Technical Memo 78556.<br />

[30] Harrington, J., R.P. Lebeau, K.A. Backes, and T.E. Dowling, 1994. Dynamic<br />

response of Jupiter’s atmosphere to the impact of comet Shoemaker-Levy 9.<br />

Nature 368, 525–527.<br />

[31] Heymsfield, A.J., and L. J. Donner, 1990. A scheme for parameterizing ice cloud<br />

water content in general circulation models. J. Atmos. Sci. 47, 1865–1877.<br />

[32] Hong, S.-Y., J. Dudhia, S.-H. Chen, 2004. A revised approach to ice microphysical<br />

processes for the bulk parameterization of clouds and precipitation. Mon.<br />

Wea. Rev. 132, 103–120.<br />

92


[33] Houze, Jr., R.A., P.V. Hobbs, P.H. Herzegh, and D.B. Parsons, 1979. Size distributions<br />

of precipitation particles in frontal clouds. J. Atmos. Sci. 36, 156–162.<br />

[34] Hueso, R. and A. Sánchez-Lavega, 2001. A three-dimensional model of moist<br />

convection for the giant planets: the Jupiter case. Icarus 151, 257–274.<br />

[35] Ingersoll, A.P., Gierasch, P.J., D. Banfield, A.R. Vasavada, and the Galileo Imaging<br />

Team, 2000. Moist convection as an energy source for the large-scale motions<br />

in Jupiter’s atmosphere. Nature 403, 630–632.<br />

[36] Ingersoll, A.P., T.E. Dowling, P.J. Gierasch, G.S. Orton, P.L. Read, A. Sánchez-<br />

Lavega, A.P. Showman, A.A. Simon-Miller, A.R. Vasavada, 2004. Dynamics of<br />

Jupiter’s Atmosphere, In: Bagenal, F., T.E. Dowling, W.B. Mckinnon (Eds.).<br />

Jupiter: The planet, Satellites and Magnetosphere, Cambridge Univ. Press, Cambridge,<br />

pp. 105–128.<br />

[37] James, E.P., O.B. Toon, and G. Schubert, 1997. A numerical microphysical model<br />

of the condensational Venus cloud. Icarus 129, 147–171.<br />

[38] Kessler, E. 1969. On the distribution and continuity of water substance in <strong>atmospheric</strong><br />

calculations. Meteor. Monogr., No.32, Amer. Meteor. Soc., 84 pp.<br />

[39] Konor, C.S. and A. Arakawa, 1997. Design of an <strong>atmospheric</strong> model based on a<br />

generalized vertical coordinate. Mon. Wea. Rev. 125, 1649–1673.<br />

[40] LeBeau, R.P. and T.E. Dowling, 1998. <strong>EPIC</strong> simulations of time-dependent,<br />

three-dimensional vortices with application to Neptune’s Great Dark Spot. Icarus<br />

132, 239–265.<br />

93


[41] Lewis, J.S. 1969. The clouds of Jupiter and the NH3-H2O and NH3-H2S systems.<br />

Icarus 10, 365–378.<br />

[42] Lin, Y.-L., R.D. Farley, and H.D. Orville, 1983. Bulk parametrization of the snow<br />

field in a cloud model. J. Climate Appl. Meteor. 22, 1065–1092.<br />

[43] Lindal, G.F., G.E. Wood, G.S. Levy, J.D. Anderson, D.N. Sweetnam, H.B. Hotz,<br />

B.J. Buckles, D.P. Holmes, P.E. Downs, V.R. Eshelman, G.L. Tyler, and T.A.<br />

Croft, 1981. The atmosphere of Jupiter: an analysis of the Voyager radio occultation<br />

measurements. J.Geophys.Res. 86, 8721–8727.<br />

[44] Limaye, S.S., 1986. Jupiter: new estimates of the mean zonal flow at the cloud<br />

level. Icarus 65, 335-352.<br />

[45] Little, B., C.D. Anger, A.P. Ingersoll, A.R. Vasavada, D.A. Senske, H.H. Breneman,<br />

W.J. Borucki, and the Galileo SSI Team, 1999. Galileo images of Lightning<br />

on Jupiter. Icarus 142, 306–323.<br />

[46] Locatelli, J.D., and P.V. Hobbs, 1974. Fall speeds and masses of solid precipitation<br />

particles. J. Geophys. Res. 79, 2185–2197.<br />

[47] Lodders, K., and B. Fegley, Jr., 1998. The Planetary scientist’s companion. Oxford<br />

University Press, 371 pp.<br />

[48] Lorenz, R.D. 1993. The Life, Death and Afterlife of a Raindrop on Titan. Planet.<br />

Space Sci. 41, 647-655.<br />

[49] Lunine, J.I., and D.A.Hunten, 1987. Moist convection and the abundance of<br />

water in the troposphere of Jupiter. Icarus 69, 566–570.<br />

94


[50] Marshall, J.S., and W. McK. Palmer, 1948. The distribution of raindrops with<br />

size. J. Meteor.5, 165–166.<br />

[51] Morrison, H., M.D. Shupe, and J. A. Curry, 2003. <strong>Model</strong>ing clouds observed at<br />

SHEBA using a bulk microphysics parameterization implemented into a singlecolumn<br />

model. J. Geophys. Res. 108, 10.1029/2002JD002229<br />

[52] Ooyama, K. V., 2001. A dynamic and thermodynamic foundation for <strong>modeling</strong><br />

the moist atmosphere with parameterized microphysics. J. Atmos. Sci. 58, 2073–<br />

2102.<br />

[53] Orton, G.S., and 16 coauthors, 1998. Characteristics of the Galileo probe entry<br />

site from Earth-based remote sensing observations. J. Geophys. Res. 103, 22,791–<br />

22,814.<br />

[54] Ortiz, J.L., G.S. Orton, A.J. Friedson, S.T. Stewart, B.M. Fisher, and J.R.<br />

Spencer, 1998. Evolution and persistence of 5-µm hot spots at the Galileo probe<br />

entry latitude. J. Geophys. Res. 103, 23,051–23,069.<br />

[55] Perry, R.H., and D.W. Green, 1997. Perry’s chemical engineers’ handbook, 7th<br />

ed. McGraw-Hill, 2640 pp.<br />

[56] Pruppacher, H.R., and J.D. Klett, 1997. Microphysics of Clouds and Precipitation.<br />

2nd ed. Kluwer Academic Publishers, 954 pp.<br />

[57] Randall, D.A., and D.G. Cripe, 1999. Alternative methods for specification of<br />

observed forcing in single-column models and cloud system models. J. Geophys.<br />

Res. 104, 24,527–24,546.<br />

95


[58] Richardson, M.I., R.J. Wilson, and A.V. Rodin, 2002. Water ice clouds in the<br />

Martian atmosphere: General circulation model experiments with a simple cloud<br />

scheme. J.Geophys.Res. 107, 1-29.<br />

[59] Rogers, R.R., and M.K. Yau, 1989. A Short Course in Cloud Physics. 3rd ed.<br />

Butterworth-Heinemann, 290 pp.<br />

[60] Rossow, W.B. 1978. Cloud Microphysics: Analysis of the clouds of Earth, Venus,<br />

Mars, and Jupiter. Icarus 36, 1–50.<br />

[61] Rotstayn, L.D., B.F. Ryan, and J.J. Katzfey, 2000. A scheme for calculation of<br />

the cloud liquid fraction in mixed-phase stratiform clouds in large-scale models.<br />

Mon. Wea. Rev. 128, 1070–1088.<br />

[62] Rutledge, S.A., and P.V.Hobbs, 1983. The mesoscale and microscale structure<br />

and organization of clouds and precipitation in mid-latitude cyclones. VIII: A<br />

model for the ”seeder-feeder” process in warm-frontal rainbands. J. Atmos. Sci.<br />

40, 1185–1206.<br />

[63] Ryan, B.F., 1996. On the global variation of precipitating layer clouds. Bull.<br />

Amer. Meteor. Soc. 77 53–70.<br />

[64] Ryan, B.F., 2000. A bulk parameterization of the ice particle size distribution<br />

and the optical properties in ice clouds. J. Atmos. Sci. 57, 1436–1451.<br />

[65] Seiff, A. 2000. Dynamics of Jupiter’s atmosphere. Nature 403, 603–605.<br />

[66] Showman, A.P. and T.E. Dowling, 2000. nonlinear simulations of Jupiter’s 5-<br />

micron hot spots. Science 289, 1737–1740.<br />

96


[67] Smith, B.A., L.A. Soderblom, T.V. Johnson, A.P. Ingersoll, S.A. Collins, E.M.<br />

Shoemaker, G.E. Hunt, H. Masursky, M.H. Carr, M.E. Davies, A.F. Cook, J.M.<br />

Boyce, G.E. Danielson, T.C. Owen, C. Sagan, R.F. Beebe, J. Veverka, R.G.<br />

Strom, J.F. McCauley, D. Morrison, G.A. Briggs, and V.E. Suomi, 1979. The<br />

Jupiter System Through the Eyes of Voyager 1. Science 204, 951-972.<br />

[68] Srivastava, R.C., 1967. A study of the effects of precipitation on cumulus <strong>dynamics</strong>.<br />

J. Atmos. Sci. 24 36–45.<br />

[69] Sromovsky, L.A., A.D. Collard, P.M. Fry, G.S. Orton, M.T. Lemmon, M.G.<br />

Tomasko, and R.S. Freedman, 1998. Galileo Probe Measurements of Thermal<br />

and Solar Radiation Fluxes in the Jovian Atmosphere. J. Geophys. Res. 103,<br />

22,929–22,977.<br />

[70] Stoker, C.R. 1986. Moist convection: A mechanism for producing the vertical<br />

structure of the Jovian equatorial plumes. Icarus 67, 106–125.<br />

[71] Stratman, P.W., A.P. Showman, T.E. Dowling, and L.A. Sromovsky, 2001. <strong>EPIC</strong><br />

simulations of bright companions to Neptune’s Great Dark Spots. Icarus 151,<br />

275–285.<br />

[72] Taylor F.W., S.K. Atreya, Th. Enccrenaz, D.M. Hunten, P.G.J. Irwin, T.C.<br />

Owen, 2004. Composition of Atmosphere of Jupiter, In: Bagenal, F., T.E. Dowling,<br />

W.B. Mckinnon (Eds.). Jupiter: The planet, Satellites and Magnetosphere,<br />

Cambridge Univ. Press, Cambridge, pp. 59–78.<br />

[73] Teixeira, M.A.C., and P.M.A. Miranda, 1997. The introduction of warm rain<br />

microphysics in the NH3D mesoscale model. CGUL Technical Report N.4, 23pp.<br />

97


[74] Thorpe, A.D., and B.J. Mason, 1966. The evaporation of ice spheres and ice<br />

crystals Brit. J. Appl. Phys. 17, 541–548.<br />

[75] Yair, Y., Z. Levin, and S. Tzivion, 1995. Microphysical processes and <strong>dynamics</strong><br />

of a Jovian thundercloud. Icarus 114, 278–299.<br />

[76] Weidenschilling, S.J., and J.S. Lewis, 1973. <strong>Atmospheric</strong> and cloud structures of<br />

the Jovian planets. Icarus 20, 465–476.<br />

[77] Weinstein, A.I., 1970. A numerical model of cumulus <strong>dynamics</strong> and microphysics.<br />

J. Atmos. Sci. 27, 246–255.<br />

[78] Wong, M.H., P.R. Mahaffy, S.K. Atreya, H.B. niemann, and T.C. Owen, 2004.<br />

Updated Galileo probe mass spectrometer measurements of carbon, oxygen, nitrogen,<br />

and sulfur on Jupiter. Icarus 171, 153–170.<br />

98


CURRICULUM VITAE<br />

NAME:<br />

Csaba Palotai<br />

ADDRESS:<br />

Department of Mechanical Engineering<br />

University of Louisville<br />

EDUCATION<br />

& TRAINING:<br />

Louisville, KY 40292<br />

M.S. Mechanical Engineering<br />

Technical University of Budapest, Hungary, 1992-2001<br />

Future Professors Program<br />

University of Louisville, 2002-2003<br />

NASA Summer School for High Performance<br />

Computational Earth and Space Sciences<br />

RESEARCH<br />

INTEREST:<br />

NASA Goddard Space Flight Center, July 12-30, 2004<br />

<strong>Atmospheric</strong> <strong>dynamics</strong> and Cloud Physics<br />

Comparative Planetology<br />

Computational Heat Transfer and Fluid Dynamics<br />

99

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!