Cloudy - Cornell Universityhosting.astro.cornell.edu/.../2015/06/cloudposter.pdf · 2017-06-01 ·...

1
a a Cloudy Preliminary results on the addition of self consistant clouds to a coupled climate and photochemistry model with a c h ance of h i g h u n certainty Abstract Here on Earth the shade of a cloud can provide momentary relief on a brutally hot summer day but for life on exoplanets clouds can mean the difference between life and death. In this work we explore the role water clouds play on the habitability of recently discovered exoplanets on the inner edge of the classically dened habitable zone and how the presence of clouds would inuence the detectability of potential biosignatures. Using a coupled climate and photochemistry model we rst simulate cloud evolution in these environments to determine surface conditions. Through the modeling process we also obtain the photochemically stable proles of various molecules that may be present in the atmosphere. These proles can be used to generate spectra that mimic future observations using reected light or transmission. Clouds can hinder our ability to identify the atmospheric composition of exoplanets and also play a big roll in the energy balance of the climate. Our work will determine the extent of these effects, provide constraints on the range of habitable environments one would expect to nd, and point to which ones would most likely be detectable in the near future. This type of study is very useful in the current stages of exoplanet characterization as the dif culty of these measurements makes fully thought out candidate prioritization an essential component of nding life sooner rather than later. Jack Madden Cornell University Carl Sagan Institute [email protected] In collaboration with Lisa Kaltenegger, and Ramses Ramirez Cloud Model No Clouds Simple Clouds Our Clouds The 1D cloud model we have implemented self consistently calculates the height and density of the liquid and ice cloud layers based on Earth observations. The cloud model convergence tests shown to the left tell us that the model has greater accuracy than simpler models that use parameterized clouds. For a given temperature and pressure there is a maximum amount of water that can be held as a vapor in the air before it starts to condense. The bottom of the cloud is determined when the humidity matches this saturation level. As warm air rises and cools more vapor condenses out. Number and size of droplets increase with height. The top of the cloud is calculated when the velocity between drops increases to the point where collisions become ef cient. The top of the cloud is calculated when the velocity between drops increases to the point where collisions become ef cient. Cloud formation depends on pressure, temperature, relative humidity, and number density of aerosols. They usually form with an updraft of warm humid air. How our clouds work a side: the surface temperature and albedo as a function of layer number u C S I Findings Conclusions It is clear that our models can be used to explore trends in how clouds interact with the other parameters in the simulation but there are many areas of uncertainty that need to be addressed. This points out the larger issue of how to accurately generate self consistent clouds in a 1D climate and photochemistry model. These trial runs have shown that clouds can have profound effects on the habitability and stability of exoplanets modeled with our code. With a model that allows for the manipulation of over 40 free parameters we chose to start with effective ux, relative humidity, and precipitation ef ciency. The other parameters were set to match Earth values as closely as possible to achieve a balance between accuracy and speed. To indirectly assess habitability we plot the surface temperature as a function of the changing parameters. Each dot represents a single run of our model with the colored dots showing a converged surface temperature and the gray dots showing where our model did not nd a stable solution. References can be found online at astro.cornell.edu/~jmadden Plots from Zsom 2012 0. 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1. 0. 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.95 1. Cloud Fraction Relative Humidity 275 280 285 290 Surface Temp. ( K) 0. 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1. 0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.95 1. 1.05 1.1 1.15 1.2 1.25 1.3 1.35 1.4 1.45 1.5 Cloud Fraction Flux fraction 240 260 280 300 Surface Temp. ( K) 0. 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1. 0. 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.95 1. Cloud Fraction Precipitation Efficiency 275.0 277.5 280.0 282.5 285.0 287.5 Surface Temp. ( K)

Transcript of Cloudy - Cornell Universityhosting.astro.cornell.edu/.../2015/06/cloudposter.pdf · 2017-06-01 ·...

Page 1: Cloudy - Cornell Universityhosting.astro.cornell.edu/.../2015/06/cloudposter.pdf · 2017-06-01 · f e t r a b e m3t d b) 0 6-n z 0 a e n z . 5s r n z e e) e n 5 o-t-y. l y) e, e

too expensive. We change the pressure grid instead. The ratio ofgrid spacing at the top to the grid spacing at the bottom of the

all possible combinations of the free parameters, but we feel thatour understanding is better served by varying only one parameter

ba

Fig. 7. Left side: the surface temperature and albedo as a function of layer number used for the convergence test of the microphysical cloud model. We assume a surfacerelative humidity of 77%, an aerosol number density of 100 cm#3, a liquid water cloud fraction of 40%, cirrus cloud fraction of 25%, and a precipitation eff ciency of 0.8. Rightside: theproperties of the cloud layers. They axis shows theheight of the layer, the color represents the radius of thedroplets in the layers. The red contours show the liquidwater cloud properties, the blue contours correspond to the ice cloud. (For interpretation of the references to color in this f gure legend, the reader is referred to the webversion of this article.)

610 A. Zsom et al./ Icarus 221 (2012) 603–616

clear-sky albedo converges as the number of vertical layers in-creases. The surface temperature difference between nz ¼570and 600 runs is 0.01 K, the surface albedos are basically equal (adifference occurs only at the f fth decimal point). Therefore weconclude that convergence is reached for nz ¼600.

Fig. 5b shows the temperature, ozone and water mixing ratiosas a function of height after convergence was reached fornz ¼600. The dotted lines show the US Standard Atmosphere(1976) prof les. The agreement between the standard atmosphereprof les and the derived prof les are not perfect in Fig. 5 for tworeasons. (1) The standard atmosphere prof les are based on glob-ally averaged measurements, therefore these prof les might notrepresent an atmosphere that is in radiative and chemical equilib-rium. (2) The correct Earth surface temperature is inf uenced byclouds, which is not accounted for in this model atmosphere.

4.2.2. Parameterized cloud modelWeadopt ahomogenouscloudmodel similar to theoneused by

Goldblatt et al. (2011) for their low level water clouds.Weassumea 40%cloud coverage, cloud deck at 0.85 bar, cloud top at 0.7 bar,mono-disperse droplets meaning that all droplets have the sameradius, 11 l m, and column density of cloud water (or water path)40g/m2. The number of layers are varied between 150 and 600with a step of 30. We found that using less than 150 layers intro-duce larger uncertainties in the surface temperature and albedobecause there are only one or two cloudy layers, which can poten-tially extend to amuchwider or narrower pressurerangethan 0.85

and 0.7 bar. The number of cloudy layers are three and twelve for150 and 600 vertical layers, respectively.

The surface temperature and albedo are shown in Fig. 6a, thetemperature, ozone and water mixing ratios are shown in Fig. 6b.The surface temperature is between 273 and 285K, the albedo ofthe planet is between 0.25 and 0.27 depending on the number oflayers used. Convergence is observed for the surface temperatureand for the planetary albedo. The cloud deck and top are not ex-actly at the prescribed pressures due to the vertical discretizationof thepressuregrid. As theexact pressure values of the cloud deckand top f uctuates, so does thesurfacetemperatureand albedo val-ues as seen in Fig. 6.

4.2.3. Themicrophysical cloud modelIn this convergence test, weuse themicrophysical cloud model

as described in Sections 3.3 and 3.4. The free parameters of themicrophysical cloud model are the number density of aerosols(nsurf ), the cloud fractions (fl, and fi), the relative humidity on thesurface (vsurf), and the precipitation eff ciency (ep). Here we usensurf ¼100cm" 3 (Miles et al., 2000), vsurf ¼0:77, which both areaverage measured Earth quantities. Finally, ep ¼0:8; f l ¼0:4, andfi ¼0:25 areused to obtain surface temperaturesequal to themea-sured average Earth value of 289K.

In apressuregridwhich is uniformly spaced in logpressure(seeSection 4.1), there are only two cloudy layers even for 600 gridcells, thus we found no real convergence. One could increase thenumber of layers above 600, but that renders the computations

ab

Fig. 5. Left side: convergence test for theclear-sky case. Thesolid lineshows thesurface temperatureas a function of thenumber of layers used in themodel. Thedotted lineshows thesurfacealbedo. Right side: the temperature, ozoneand water mixing ratios as a function of height for nz ¼600. Thedotted lines show theUSStandard Atmosphere(1976) prof les.

a b

A. Zsom et al./ Icarus 221 (2012) 603–616 609

CloudyPreliminary results on the addition of self consistant

clouds to a coupled climate and photochemistry model

with a chance of high uncertainty

Abstract Here on Earth the shade of a cloud can provide momentary relief on a brutally hot summer day but for life on exoplanets clouds can mean the difference between life and death. In this work we explore the role water clouds play on the habitability of recently discovered exoplanets on the inner edge of the classically defined habitable zone and how the presence of clouds would influence the detectability of potential biosignatures. Using a coupled climate and photochemistry model we first simulate cloud evolution in these environments to determine surface conditions. Through the modeling process we also obtain the photochemically stable profiles of various molecules that may be present in the atmosphere. These profiles can be used to generate spectra that mimic future observations using reflected light or transmission. Clouds can hinder our ability to identify the atmospheric composition of exoplanets and also play a big roll in the energy balance of the climate. Our work will determine the extent of these effects, provide constraints on the range of habitable environments one would expect to find, and point to which ones would most likely be detectable in the near future. This type of study is very useful in the current stages of exoplanet characterization as the difficulty of these measurements makes fully thought out candidate prioritization an essential component of finding life sooner rather than later.

Jack MaddenCornell UniversityCarl Sagan [email protected]

In collaboration with Lisa Kaltenegger, and Ramses Ramirez

Cloud Model

NoClouds

SimpleClouds

OurClouds

The 1D cloud model we have implemented self consistently calculates the height and density of the liquid and ice cloud layers based on Earth observations. The cloud model convergence tests shown to the left tell us that the model has greater accuracy than simpler models that use parameterized clouds.

For a given temperature and pressure there is a maximum amount of water that can be held as a vapor in the air before it starts to condense. The bottom of the cloud is determined when the humidity matches this saturation level.

As warm air rises and cools more vapor condenses out. Number and size of droplets increase with height.

The top of the cloud is calculated when the velocity between drops

increases to the point where collisions become efficient.

The top of the cloud is calculated when the velocity between drops

increases to the point where collisions become efficient.

Cloud formation depends on pressure, temperature, relative humidity, and number density of aerosols. They usually form with an updraft of warm humid air.

How our clouds work

clear-sky albedo converges as the number of vertical layers in-creases. The surface temperature difference between nz ¼570and 600 runs is 0.01 K, the surface albedos are basically equal (adifference occurs only at the f fth decimal point). Therefore weconclude that convergence is reached for nz ¼600.

Fig. 5b shows the temperature, ozone and water mixing ratiosas a function of height after convergence was reached fornz ¼600. The dotted lines show the US Standard Atmosphere(1976) prof les. The agreement between the standard atmosphereprof les and the derived prof les are not perfect in Fig. 5 for tworeasons. (1) The standard atmosphere prof les are based on glob-ally averaged measurements, therefore these prof les might notrepresent an atmosphere that is in radiative and chemical equilib-rium. (2) The correct Earth surface temperature is inf uenced byclouds, which is not accounted for in this model atmosphere.

4.2.2. Parameterized cloud modelWeadopt ahomogenouscloudmodel similar to theoneused by

Goldblatt et al. (2011) for their low level water clouds.Weassumea 40%cloud coverage, cloud deck at 0.85 bar, cloud top at 0.7 bar,mono-disperse droplets meaning that all droplets have the sameradius, 11 l m, and column density of cloud water (or water path)40g/m2. The number of layers are varied between 150 and 600with a step of 30. We found that using less than 150 layers intro-duce larger uncertainties in the surface temperature and albedobecause there are only one or two cloudy layers, which can poten-tially extend to amuchwider or narrower pressurerangethan 0.85

and 0.7 bar. The number of cloudy layers are three and twelve for150 and 600 vertical layers, respectively.

The surface temperature and albedo are shown in Fig. 6a, thetemperature, ozone and water mixing ratios are shown in Fig. 6b.The surface temperature is between 273 and 285K, the albedo ofthe planet is between 0.25 and 0.27 depending on the number oflayers used. Convergence is observed for the surface temperatureand for the planetary albedo. The cloud deck and top are not ex-actly at the prescribed pressures due to the vertical discretizationof thepressuregrid. As theexact pressure values of the cloud deckand top f uctuates, so does thesurfacetemperatureand albedo val-ues as seen in Fig. 6.

4.2.3. Themicrophysical cloud modelIn this convergence test, weuse themicrophysical cloud model

as described in Sections 3.3 and 3.4. The free parameters of themicrophysical cloud model are the number density of aerosols(nsurf ), the cloud fractions (fl, and fi), the relative humidity on thesurface (vsurf), and the precipitation eff ciency (ep). Here we usensurf ¼100cm" 3 (Miles et al., 2000), vsurf ¼0:77, which both areaverage measured Earth quantities. Finally, ep ¼0:8; f l ¼0:4, andfi ¼0:25 areused to obtain surface temperaturesequal to themea-sured average Earth value of 289K.

In apressuregridwhich is uniformly spaced in logpressure(seeSection 4.1), there are only two cloudy layers even for 600 gridcells, thus we found no real convergence. One could increase thenumber of layers above 600, but that renders the computations

ab

Fig. 5. Left side: convergence test for theclear-sky case. Thesolid lineshows thesurface temperatureas a function of thenumber of layers used in themodel. Thedotted lineshows thesurfacealbedo. Right side: the temperature, ozoneand water mixing ratios as a function of height for nz ¼600. Thedotted lines show theUSStandard Atmosphere(1976) prof les.

a b

Fig. 6. Left side: the surface temperature and albedo as a function of layer number used for theparameterized cloud convergence test. Weassumea40%cloud coverage, thecloud deck and top are located at P ¼0:85and 0.7bars, respectively, thedropletshavearadiusof 11 l m,and thetotal water path is40g/m2 (Goldblatt et al.,2011).Right side:the temperature, water and ozonemixing ratio prof les as a function of height for nz ¼600. The dotted lines show the US Standard Atmosphere (1976) prof les.

A. Zsom et al./ Icarus 221 (2012) 603–616 609

CSI

Findings

ConclusionsIt is clear that our models can be used to explore trends in how clouds interact with the other parameters in the simulation but there are many areas of uncertainty that need to be addressed. This points out the larger issue of how to accurately generate self consistent clouds in a 1D climate and photochemistry model. These trial runs have shown that clouds can have profound effects on the habitability and stability of exoplanets modeled with our code.

With a model that allows for the manipulation of over 40 free parameters we chose to start with effective flux, relative humidity, and precipitation efficiency. The other parameters were set to match Earth values as closely as possible to achieve a balance between accuracy and speed. To indirectly assess habitability we plot the surface temperature as a function of the changing parameters. Each dot represents a single run of our model with the colored dots showing a converged surface temperature and the gray dots showing where our model did not find a stable solution.

References can be found online at astro.cornell.edu/~jmadden

Plots from Zsom 2012

0. 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.

0.0.05

0.10.15

0.20.25

0.30.35

0.40.45

0.50.55

0.60.65

0.70.75

0.80.85

0.90.95

1.

Cloud Fraction

Rela

tive

Hum

idity

275

280

285

290

SurfaceTem

p.(K)

0. 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.

0.50.55

0.60.65

0.70.75

0.80.85

0.90.95

1.1.05

1.11.15

1.21.25

1.31.35

1.41.45

1.5

Cloud Fraction

Flux

fract

ion

240

260

280

300

SurfaceTem

p.(K)

0. 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.

0.0.05

0.10.15

0.20.25

0.30.35

0.40.45

0.50.55

0.60.65

0.70.75

0.80.85

0.90.95

1.

Cloud Fraction

Prec

ipita

tion

Effic

ienc

y

275.0

277.5

280.0

282.5

285.0

287.5

SurfaceTem

p.(K)