The Four Elements: A Quest for Atmosphere (continued)

This post is for those, who will be modeling their planet atmospheres in detail. If you have used Starform (accrete) code to generate your system and evolved the system with Mercury afterwards, using a climate model for your planet would be a good choice to continue.

Before I will go into details with data preparation and actual modeling, there are multiple issues with compiling some parts of the FORTRAN code, especially the ones containing f90 or fIV/f66 features. For *nix systems ifort and f90 compilers are the necessary choice.

Now, I know for sure f77, gfortran, g95 or FTN95 won’t be much help. (The latter is Windows only; now with this one I’m not exactly sure, but I couldn’t make it compile with 8 byte-integer AND real options, it allowed me one or another, and not both; even if it is possible, it is NOT OBVIOUS HOW at all.) For Windows users issues are resolved with Compaq Visual FORTRAN Professional v6.6, for example, or Intel’s ifort compiler if you have one. I like Compaq better because you don’t need to install Visual Studio additionally, and it handles pretty much everything. Please note that version 6.5 or lower won’t do because of the integer/real intrinsic data types support (8 bytes are necessary).

Once you are technically set, you can move onto the next phase of atmosphere/climate modeling – data preparation. It is also assumed that you have decided on the atmospheric composition. I’m not an astrophysicist, so everything I do below this point is what I have came up so far by guess, trial and error method (yes, I’m THAT persistent) and is not claimed to be flawless.

Comments, suggestions and corrections are welcome.

Data preparation

I’ll start with the 1-D coupled model and there are several things to do before actual modeling.

Stellar spectrum. If you are using BaSeL server, your output is the flux moment given in erg/cm^2/s/Hz/sr. To get this in ergs/s/cm^2/Hz you need to multiply it by PI to get rid of steradians (sr). Then, flux Flambda, in units ergs/s/cm^2/Angstrom, can be calculated using the equation Flambda=0.4*flux_moment*c/Lambda^2, where c=2.997925e17 is the velocity of light and Lambda is the wavelength in nanometers. The numerical factor of 0.4 (=4*0.1) in equation above comes from the conversion of the flux moment into flux (*4) and from the conversion of flux per nanometer into flux per Angstrom (*0.1). However Flambda still is a flux at the surface of a star.

The flux at the top of the atmosphere is used in models; the altitude of 100 km is taken as „the top”. For Earth, an altitude of 120 km marks the boundary where atmospheric effects become noticeable during spacecraft re-entry. Most of the atmosphere (99.9999 percent) is below 100 km, although in the rarefied region above this there are auroras and other atmospheric effects.

Now, flux, S, from a star drops off with increasing distance. In fact, it decreases with the square of the radial distance, r, from the star, as proportional to 1/r^2. You will need to provide distance in parsecs (see Stellar distance d in pc below).

Stellar flux with distance
Stellar flux with distance

There is also an option to obtain stellar spectrum by using ATLAS9 (or ATLAS12) and SYNTHE programs by Kurucz and Castelli (or SPECTRUM by Richard O. Gray, needs Atlas or MARCS stellar atmosphere model anyway). Here you can be more detailed with your star, providing additional initial conditions like microturbulent velocity* and mixing length parameter** into model. Then, there is also Chris Sneden’s MOOG.

*Microtubulent velocity. This is one of the fundamental stellar properties, a standard parameter in one-dimensional analyses of solar-type stars. For ATLAS models the value of microturbulent velocity, ξ, can be chosen [among 0, 1, 2, 4, etc. km/s] appropriate for the star’s surface gravity between the two velocities that bracket the microturbulent velocity. HOWEVER, the relation given by Kirby (Eq. 2 of Kirby et al. 2009, also here): ξ (km/s) = (2.13 ± 0.05) − (0.23 ± 0.03)*log g is only appropriate for giants, not sub-giants or dwarfs. Of course, for dwarfs there also exists a correlation between the luminosity class and the surface gravity (log g), the microturbulent velocity (t), and the metallicity ([M/H]) (Gray et al, 2000).

Dwarfs (V class) from F to G have the smallest microturbulent velocity, thus, technically, the higher the log g, the lower the microturbulent velocity. For precise results the technique of partial correlation can be employed, since the relationships between three or more random variables are being investigated. The trends for K stars (at least up to mid-K) are probably mainly similar to G dwarfs (Spectral studies of K dwarfs, Spectroscopic Properties of Cool Stars).

**Mixing length (ML) parameter. It is the alpha (a = l/Hp) parameter used in stellar model, which represents the ratio between the mean free path of a convective element (l) and the pressure scale height (Hp). The variations of this parameter strongly affect the structure of the outer envelope (i.e. radius and temperature). In fact, this parameter determines the efficiency of energy transport by convection in the outermost layer of a star: for a given stellar luminosity, it fixes the radius of the star, hence its temperature and color. In case of a real star, evolutionary tracks must then be calibrated by comparison with stellar radii and/or temperatures derived from observations. The most obvious ML calibrator is the Sun. The ML parameter can be fixed by constraining theoretical solar models (i.e. with solar mass, age and chemical composition) to reproduce the solar radius. In low-mass stellar models the derived stellar radii depend on the opacity which significantly contributes in determining the temperature gradient in their turbulent external layers. Hence, stellar models, based on different opacity tables, could require different values of alpha. It means that a homogeneous dataset of stellar temperatures at different metallicities to properly calibrate the ML parameter is urgently needed before any further attempt to use evolutionary models to derive relevant properties of stellar populations (Ferraro et al. 2006).

Lyman alpha flux at the planet (xLy). Lyman-alpha line is the brightest emission line of neutral hydrogen at the wavelength of 1215.67 A (121.567 nm) in the spectrum (in some papers 1215.7 A or 1216 A is mentioned). Stellar Lya emission lines are important spectral features in the context of exoplanet stellar environment and stellar physics. The Lya line is used as a proxy for determining the temperature and pressure profiles of upper stellar atmospheres. Lya flux is extremely variable with time. To a first approximation the solar La flux is composed of a quiet and of an active component. The Sun’s active component changes with the 27 days period (1 rotation around its axis); the quiet one with the 11 year solar cycle. For the Sun, the integrated Lya line flux may change by 37% during one rotation and up to 50% over a couple of years.

These fluxes can be obtained from the catalogue; in case of a synthetic star Lyman-alpha flux is the flux integrated over the whole Lya emission line of the synthetic spectrum. The easier way around is to compute a synthetic stellar Lya profile of a given linewidth, applying a simple linear rescaling in wavelength to the solar Lya profile. Real stellar profiles may be poorly represented by such a simple rescaling of the solar profile.

Math Box 1 – Some interesting data about your star

If you made yourself a synthetic star, some things are not given, but can be obtained through calculation. Now I was thinking if I need to include this in one of my posts, but heck, maybe it can be useful to someone.

If you know B-V color (from synthetic spectrum or from stellar data), then use it. Otherwise it can be calculated as B-V = (-3.684*LOG(Teff,K))+14.555. You can compare this data with the one from BaSeL result, for example, and make necessary adjustments.

Stellar rotation (vsin i)

Now, this one is tied to B-V and stellar age. I picked up formulas from (Barnes, 2007). Gyrochronology permits the derivation of ages for solar- and late-type main sequence stars using only their rotation periods and colors. If you know the age of your star and B-V color, you can go the other way around.

f(B-V) =0.7725*(((B-V)-0.4)^0.601), where B-V is the color;

g(t) = T^0.5119, where T is star age in Myr;

Rotational period P, days, is found as P = g(t)*f(B-V);

And finally, stellar equatorial rotational velocity, vsini, km/s, is found from the rotational period:

vsini = ((2*PI()*(R*2))/P)*(1/86400), where R is stellar radius and P is rotational period in days.

Stellar activity cycle

Approximate stellar cycle P, yrs, can be calculated as P = -1.22+(14.14*(B-V)). This is how long it takes your star to go from the minimum to maximum stellar activity. The complete cycle is, obviously, double of that.

Stellar distance d in pc. For your synthetic star you can take one of the provided distances (and manually scale the spectrum) for comparison with the VPL spectral data (I used 3.2 parsecs). If you used a real star, the real distance from Earth must be provided. It is needed for conversion to values expected for an Earth-like planet (see The Afac correction parameter). The flux will be scaled according to this distance.

The Afac correction parameter. To convert to values expected for an Earth-like planet, the measured UV fluxes were multiplied by ((206265*d )^2)*(Lsun/Lstar)*Afac, here d is the distance in parsecs (1 pc = 206264.806 =206205 AU), Lstar and Lsun are the respective bolometric luminosities of the star and the Sun, and Afac is a correction factor that accounts for the change in the planet’s albedo with the wavelength of the incident radiation. Values of Afac of 1.11 for the F2V star and 0.95 for the K2V star were obtained by scaling the “water loss” Seff limits in Habitable Zones around Main Sequence Stars“, Kasting, J.F., Whitmire, D.P. & Reynolds, R.T. Icarus 101, 108-128 (1993), Table III, by the effective radiating temperature of the stars, using a quadratic fit to the listed temperatures. This procedure ensures that hypothetical planets would have the same surface temperature as the Earth (~288 K) if other climatic factors (e.g. cloudiness and greenhouse gas concentrations) are the same.

I went the other way around scaling given Afac numbers (x) to stellar temperatures (y):

x y notes
0.9 3450 (M3.5V, AD Leo)
0.95 5084 (K2V, ε Eridani)
1 5778 (G2V, Sun)
1.11 6930 (F2V, σ Boo)

Using quadratic regression I got coefficients for equation y=ax^2+bx+c:

a = -70906.1520; b = 158654.0223; c = -86698.5002

For any given temperature (K) within the range of [3450; 6930] you can solve the equation to find the roots:

x1 = (-b+SQRT((b^2)-4ac))/2a; x2 = (-b-SQRT((b^2)-4ac))/2a

Use the root that fits the range of Afac [0,9; 1,11] – in all cases here it will be x1.

Height of the tropopause. In planet.dat file you will need to provide the height of the tropopause. It is found to be strongly sensitive to the temperature at the planet’s surface through changes in the moisture distribution and its resulting radiative effects. The tropopause height is less sensitive to changes in the ozone distribution and hardly sensitive at all to moderate changes in the planet’s rotation rate (Thuburn & Craig, 1996).

Coupled photochemical and radiative/convective atmosphere model. When your data is ready, you can finally make executables and run your models. Make sure that everything in the files is set as you want it. And if you are running a coupled model, make sure your folder/file tree is put like it should be, and the couple switches in the code are on. Then, compile and run.

Some more alternatives

There is a very interesting model suite called Most for atmosphere and climate. It includes Planet Simulator, PUMA (The Portable University Model of the Atmosphere) and SAM (The Shallow Atmosphere Model) along with the Graphical User Interface, the Model Starter (MoSt), the postprocessor Burn7 and all manuals.

NOTE: There is no cake for Cygwin users here. You can get Most suite running under Cygwin (make sure X11 is properly set up), but the GUI is remarkably (read: awfully) slower than on Linux. Dual boot (Win/Linux) or “virtual machine” is the salvation.

Another model I’m going to mention here is called EPIC as in Explicit Planetary Isentropic-Coordinate general circulation atmospheric model. This model is implemented in principle for all known atmospheres (one for all!); for terrestrial-planet applications the EPICwiki suggests using version 4.x. (I would love to try this one once I have free time.)

# Have fun digesting and rejoice – there will be no more modeling for now (unless I’ll run into a proper planetary interior model. Yum.)

## Also, this is probably the last post this year but if you have questions or want to discuss something, feel free to drop a line. HAPPY WINTER HOLIDAYS and come back in January for more! 😉

The Four Elements: A Quest for Atmosphere

This journal entry is dedicated to planetary atmospheres, since this is probably the most difficult part of a worldbuilder’s journey. Of course, you can have things easier way, say, if you have earthlike conditions or you have additional data laid out for you in case of a standard star. But what if things were different?

I’m going to explain each step, adding some theoretical info as well. The Four Elements series will cover topics such as the atmosphere, the ocean, geological activity, insolation and planetary climate. And maybe something else I’ll find necessary to add.

The quest doesn’t start where one might think it does – not on a planet, but on a star. So, long before we can model our planetary atmosphere or anything else related to it, first we should consider stellar spectrum. There are several ways of getting it: taking a real detailed spectrum for existing star or computing a synthetic stellar spectrum. It all depends on your star of choice and how deep you are willing to go into details.

Why is this important? To design your own exclusive and unique planetary atmosphere model: maybe the air is so thin at altitude of 5 km, so your local people will never go to the mountains, for example. Whatever it is, you’ll get the full set of necessary parameters, some of which you might have had imagined, and some of which you probably didn’t expect at all. Anyway, you will get to know your planet better.

Don’t be afraid of experimenting with things. They are not as hard or incomprehensible as might appear to be.

Planet atmosphere

The atmosphere is an envelope of gas mixture around the planet. It is held down by gravity, and the weight of that gas is pressure (as in mass times g). The total pressure is the sum of partial pressures of gasses in the mixture. The partial pressure is the contribution of a particular gas constituent to the total pressure, and is found as the total pressure times the volume fraction of gas component.

The proportion of gases found in the atmosphere changes with altitude. Distinct layers (such as troposphere, stratosphere, etc.) are identified using thermal characteristics, chemical composition, molecule movement, and density.

Individual molecules are moving freely in gas and if their motion velocity exceeds the planet’s escape velocity, the molecules will escape into space from the outer edge of the atmosphere. A certain amount will always exceed escape velocity, and if that percentage is too high, the atmosphere will leak away in a geologically short term. Thus, enough gravity is necessary to hold the atmosphere. The outer atmosphere temperature plays a vital role in this process as well, since gas molecules travel faster with increasing temperature. The hotter the exosphere is, the greater gravity must be. To keep things in balance, worlds closer to their stars must be larger to hold atmospheres equivalent to those around cooler worlds. Thus, atmospheric composition is also important, because lighter molecules move faster at the same temperature. Same surface gravity can keep one molecules, but can’t hold others; in case of Earth hydrogen and helium are too light for our gravity.

Composition and pressure are not completely free parameters, though. They are influenced and modified by chemical reactions with the surface of the planet (e.g. atmosphere interaction with crustal rocks over time in the carbonate-silicate cycle), living things and photodissociation (stellar UV light breaks up the hydrogen-bearing compounds like water, ammonia and methane) at the outer edge of the atmosphere. The atmosphere changes over geological time along with the evolution of the star, life and loss of lighter gasses.

Terrestrial-like planets may obtain atmospheres from three primary sources: capture of nebular gases, degassing during accretion, and degassing from subsequent tectonic activity. While capture of gases is vital for gas giants, low-mass terrestrial planets are unable to capture and retain nebula gases, which also may have largely dissipated from the inner solar system by the time of final planetary accretion. Atmospheric mass and composition for terrestrial planets is therefore closely related to the composition of the solid planet (Elkins-Tanton & Seager, 2008a).

In case of humans and animals the atmosphere has limits on its composition. To be breathable, it must have levels of molecular oxygen (O2) between 0.16 and 0.5 atm; higher concentrations of oxygen are toxic (severe cases can result in cell damage and death), lower than minimum are not enough to support human life. Hypoxia (oxygen deprivation) and sudden unconsciousness becomes a problem with an oxygen partial pressure of less than 0.16 atm. Hyperoxia (excess oxygen in body tissues), involving convulsions, becomes a problem when oxygen partial pressure is too high. Our present atmosphere contains 21% molecular oxygen (partial pressure of 0.21 atm).

Also, to prevent nitrogen narcosis under high pressures (the diver’s “rapture of the deep”) the partial pressure of nitrogen (N2) must be less than 3 atm.

As for other toxic stuff, the level of carbon dioxide must be less than 0.02 atm to breathe indefinitely, and less than 0.005 atm to avoid physiological stresses. In case of CO2 concentration above normal levels the only habitable places for humans might be high regions, like mountains. But then again, too little oxygen higher up can be troublesome.

Many plants, however, can survive and thrive in low oxygen-high CO2 environment. Earth’s plants will grow in many atmospheres that are unbreathable to humans and animals, unless the runaway greenhouse ruins the place completely (like Venus).

By building planet atmosphere and climate models you can see what are the boundaries for life under different stars and atmospheres. How fast the atmosphere thins upward? What are the properties of layers (altitudes, temperatures, pressures, composition, ozone layer (ozone is also toxic), gravitational pull, etc.)? What is the climate and weather pattern? The broader applications for the model include your planet aerospace or colonization/terraforming history, if applicable. The thickness of the atmosphere has some consequences. The thicker it is (and/or the lower the gravity), the easier flight is. Sound also travels better in a denser medium. Storms can be more intense if mass of moving air is greater.

More detailed description of atmospheres is beyond the scope of this article, but can be found on the Internet or in textbooks. In fact, if you know little about how atmospheres work, further reading into subject is required before building anything. Some useful book titles are listed in my LibraryThing catalogue, which is constantly growing. My goal here is to describe the tools: what data for models is required and how those models can be used to produce desirable results.

Climate dynamics model

The MITgcm (MIT General Circulation Model) is a numerical model designed for study of the atmosphere, ocean, and climate dynamics. MITgcm is freely available to all and can be run on a home pc or laptop, and is enough to play with your planet in detail.

There are some other models, such as NASA/GISS 4×3 Atmosphere-Ocean Model or the Community Earth System Model (CESM).

Running the NASA/GISS model requires a significant investment in time and money, and it is designed to run on multiprocessor machines. It can reproduce the seasonal and regional mean values and variations of climate quantities such as temperature, pressure, precipitation, cloud cover, and radiation with reasonable degrees of precision, and many other things. I do not recommend this one for our purpose (though, if you own a multiprocessor workstation, you can try).

The CESM is also designed for simulating Earth’s climate system, but, as with the NASA/GISS, it is not simple at all. It is a coupled climate model composed of five separate models simultaneously simulating atmosphere, ocean, land, land-ice, and sea-ice.

However, before you can model weather and climate, an atmospheric layered model is required. You can take earthlike model (e.g. Standard Atmosphere) or you can make your own. For the latter purpose you’ll need another piece of code, described in the section below.

Photochemical and radiative/convective atmosphere models

The coupled photochemical and radiative/convective atmosphere model was used to study earthlike planets around different types of stars: F2V, G2V (Sun), K2V (Segura et al., Astrobiology, 2003) and M stars (Segura et al., Astrobiology, 2005); Grenfell et al. 2006, 2011; Kasting et al. 1996.

This model requires stellar spectrum, which can be taken from the database or synthesized. Some spectra are hard to find. The ones used by Segura et al. can be taken from the VPL site.

Stellar flux greatly influences chemical processes in the atmosphere and biological processes on the planet. Each star has its individual flux “signature”.

In Segura’s model the “Earth” is assumed to be at a distance equivalent to 1 AU in the extrasolar planet system. The orbital radius is scaled according to stellar luminosity, and the planet is then moved inward or outward until its calculated surface temperature is 288 K. Also, the term “mixing ratio” has the same meaning as “mole fraction”.

Temperatures. Credit: Segura et al. 2003
Temperatures. Credit: Segura et al. 2003

The planet around the F star develops a thicker ozone layer because of the abundance of short-wavelength UV radiation (lambda < 200 nm) that can dissociate molecular oxygen.

Ozone number density. Credit: Segura et al. 2003
Ozone number density. Credit: Segura et al. 2003

The surface UV flux increases with decreasing partial pressure of O2, but the behavior is very nonlinear. Good UV shield develops above 10^-2 of present atmospheric level of O2.

M stars emit very little near-UV radiation (200-300 nm), but active M (and, in fact, early K) stars emit lots of UV radiation shortward of 200 nm (chromospheric emission). One can therefore split molecular oxygen (and, hence, make ozone), but the ozone photochemistry is very different. Methane in Earth’s atmosphere is mostly destroyed in chemical reactions triggered by UV-flux at 310 nm. In atmospheres near M stars the lifetime for methane is long.

Synthetic stellar spectrum

If you have a star type that is not on the VPL list of spectra, acquiring a synthetic spectrum is where you’ll have to start building your model. There are numerous ways and software packages to compute a synthetic spectrum, but the easiest one is to use the BaSeL interactive server: it saves time and sanity. This tool presents a user-friendly interface of an interpolation engine, that allows on-line computations of synthetic stellar spectra for any given set of fundamental parameters Teff, log g and [Fe/H]. More info about BaSeL is found in “The BaSeL interactive web server: a tool for stellar physics”. Please note that fundamental parameters are taken from the real star’s data or computed stellar evolution model.

# Have fun and more modeling to follow.

Setting, Worldbuilding and Geofiction

I’ve seen some debating on the Internet about Setting vs. Worldbuilding (here and here, for example). None of them provided a clear picture where the separation line should lie. But the separation line is so great, that, in fact, it is not just the line, it’s the whole damn canyon.

Setting exists only in a novel, play, film, etc. It has no purpose outside it. None. It is not the same as ‘worldbuilding’, because it is the immediate place and time of action, and the social environment of an individual, the POV character (and POVs are biased). It doesn’t go beyond that, but it sets the mood and it is what you currently see with your reader’s (viewer’s) eyes. Setting can be static or dynamic (depends on how the story is told), but it is not a process.

Worldbuilding, on the other hand, is the term describing a process of construction, engineering, if you must, of a fictional world regardless of how great/small it needs to be. It can include everything from physics to cultures, to languages and even music. It is a construction site of a ‘house’, in which each apartment can become a setting at some point in time. It’s a process of designing a framework, a ‘software package’ to write, debug, compile and build another piece of ‘software’ which is a story or what-have-you. You ‘build’ because you are structuring your ideas. You build a system. One of the great examples of such framework is the Orion’s Arm Project. It is self-sufficing and provides settings for stories. Another fine example are The Hard Return books by A. J. Klassen (AJ had published ‘book 0’ so far, but there are more on the way, and as a witness of the writing process and occasional consultant pal I can tell they’re awesome.) Love or hate it, but if you are a writer, a game designer, or anything of a sort, you are building a world.

Geofiction is almost self-explanatory. At least the ‘fiction’ part, which is exactly the statement of difference in purpose from worldbuilding. Geofiction includes everything worldbuilding does, but it exists for its own sake. A great example of geofiction is Planet Furaha.

Designing a Planetary System: Extension

Multiple star systems are not the majority in the Universe, even if in our Galaxy only 30 percent of stars are single, like our Sun. Massive stars tend to be more “family-oriented” than low-mass stars. For example, only 15 to 25 percent of M class stars are in binary or multiple systems comparing to 2/3 of G class stars, and G class makes only 7 percent of stars we see. The reason high-mass stars are often in multiples while low-mass ones are not is due to differences in how they form.

Star formation usually occurs in dense turbulent clouds of molecular Hydrogen. The natal environment influences star formation through the complex interplay of gravity, magnetic fields, and supersonic turbulence. Observations suggest that massive stars form through disk accretion in direct analogy to the formation of low-mass stars. However, several aspects distinguish high and low-mass star formation despite the broad similarity of the observed outflow and ejection phenomena.

Surprisingly, the nearby Taurus-Auriga star-forming region has a very high fraction of binaries for G stars and probably even for M stars. This produced a thought that most stars might be forming as multiples, but later these systems are broken apart.

High-mass vs. low-mass stars and the multiplet frequency

Low-mass star systems are observed to have a much lower binary fraction than higher mass stars (Lada 2006). In the recent simulations of a turbulent molecular cloud destined to form a small cluster of low-mass stars several classes of systems were produced: isolated stars, binaries and multiples formed via the fragmentation of a turbulent core, and binaries and multiples formed via the fragmentation of a disk; however, dynamical filament fragmentation is the dominant mechanism forming low-mass stars and binary systems, rather than disk fragmentation. (Offner et al. 2010). Loosely bound companions can be stripped by close encounters within the protocluster. Although previous studies suggested that multiple star formation via the fragmentation of a disk was limited to large mass ratio systems, recent work such as Stamatellos & Whitworth (2009) and Kratter et al. (2010a) has shown that when disks continue to be fed at their outer edges, the companions can grow substantially. Protostellar disks that are sufficiently massive and extended to fragment might be an important site for forming brown dwarfs and planetary mass objects (Stamatellos et al. 2011). Cores that form brown dwarfs would have to be very small and very dense to be bound (Offner et al. 2008).

Protostellar cores with a mass of a few tenths to a few hundreds of solar appear to be more turbulent, with gas within generally moving at higher velocities. Those are more prone to fragmentation, giving birth to binaries or multiple stars. Such massive star forming sites are rare and thus tend to be farther from Earth (> 400 pc) than low-mass star forming regions (around 100 pc). High-mass star formation occurs in clusters with high stellar densities. In addition, massive stars destroy their natal environment via HII regions. Their accretion disks are deeply embedded in dusty envelopes and ultra-compact HII regions become visible only after star formation is nearly complete. Observations of HII regions produced by massive stars are a prime tool for extragalactic astronomers to determine the star formation rate and abundances in galaxies.

Massive stars are shorter-lived and interact more energetically with their surrounding environment than their low-mass counterparts; they reach main sequence faster and begin nuclear burning while still embedded within and accreting from the circumstellar envelope. They also can be potential hazards to planetary formation: many low-mass stars are born in clusters containing massive stars whose UV radiation can destroy protoplanetary disks.

The temperature structure of the collapsing gas strongly affects the fragmentation of star-forming interstellar clouds and the resulting stellar initial mass function (IMF). Radiation feedback from embedded stars plays an important role in determining the IMF, since it can modify the outcome as the collapse proceeds (Krumholz et al. 2010). Radiation removes energy, allowing a collapsing cloud to maintain a nearly constant, low temperature as its density and gravitational binding energy rise by many orders of magnitude. Radiative transfer processes can be roughly broken into three categories: thermal feedback, in which collapsing gas and stars heat the gas and thereby change its pressure; force feedback, in which radiation exerts forces on the gas that alter its motion; and chemical feedback, in which radiation changes the chemical state of the gas (e.g. by ionizing it), and this chemical change affects the dynamics (Krumholz 2010).

Column densities L= 0.1, M=1.0, H=10.0 g cm-2 (left to right column). Credit: Krumholz et al. 2010
Column densities L= 0.1, M=1.0, H=10.0 g cm-2 (left to right column). Credit: Krumholz et al. 2010

In this image surface column densities L= 0.1, M=1.0, H=10.0 g cm-2 are shown. As density increases, the suppression of fragmentation increases: (L) small cluster, no massive stars, depleted disks; (M) massive binary with 2 circumstellar disks and large circumbinary disk; (H) single large disk with single massive star.

The effects of radiative heating depend strongly on the surface density of the collapsing clouds, which determines effectiveness of trapping radiation and accretion luminosities of forming stars. Surface density is also an important factor in binary and multiple star formation. Higher surface density clouds have higher accretion rates and exhibit enhanced radiative heating feedback, diminished disk fragmentation and host more massive primary stars with less massive companions (Cunningham et al. 2011).

Observations indicate most massive O-stars have one or more companions; binaries are common (> 59%) (Gies 2008). Massive protostellar disks are unstable to fragmentation at R ≥ 150AU for a star mass of 4 or more solar masses (Kratter & Matzner 2006) and cores with masses above 20 solar will form a multiple through disk fragmentation (Kratter & Matzner 2007). Radiation pressure does not limit stellar masses, but the instabilities that allow accretion to continue lead to small multiple systems (Krumholz et al. 2009).

The upper limit on the final companion star frequency in the system can be estimated (Bate 2004).

Primordial stars and their legacy

The similar pattern appears in formation of previously believed to be solitary primordial stars. Recent studies suggest that loners were rather an exception, than the rule.

First massive stars were probably accompanied by smaller stars, more similar to our Sun. Due to encounters with their neighbors some of the small stars may have been ejected from their birth group before they had grown into massive stars. This could indicate primordial stars with a broad range of masses: short-lived, high mass stars capable of enriching the cosmic gas with the first heavy chemical elements and produced first black holes that are alive and well today, and long-lived, low-mass stars which could survive for billions of years and maybe even to the present day.

Are single stars really single?

Again, similar fragmentation happens in the massive protoplanetary discs to produce gas giants, when a gas patch in protoplanetary disk collapses directly into gas giant planet (1 Jupiter mass or larger) due to gravitational instability. Gravitational instabilities can occur in any region of a gas disk that becomes sufficiently cool or develops a high enough surface density, be it a star or a planet formation site. Kratter et al. 2009 suggests that planets formed that way might be failed binaries.

In 1970 Stephen Dole performed planet accretion simulation which produced interesting results. In our Galaxy, the average separation of binary components is about 20 AU, corresponding roughly to the orbital distances of the jupiter-mass gas giants in our solar system (Jupiter and Saturn have often been called “failed stars”; however, both probably contain rocky cores, so they are definitely planets). By increasing the density of the initial protocloud an order of magnitude higher than before, Dole’s program generated larger and larger jovians. Eventually in one high-density run, a class K6 orange dwarf star appears near Saturn’s present orbit, along with two superjupiters and a faint red dwarf further sunward. No terrestrials were formed.

Planetary system accretion. A is the density parameter. Solid circles – terrestrial planets, horizontal shading – gas giants, cross-hatching designate red dwarfs, open circle represents orange dwarf. Credit: S. Dole, 1970

This study suggested that Jovians (brown dwarfs can be mentioned here too now) multiply at the expense of terrestrials. An increase of one critical parameter – the nebular density – resulted in the generation of binary and multiple star systems, and close companionship might lead to eventual exclusion of terrestrial worlds.

But things appear to be more complicated.

Terrestrial planets in multiple star systems

Multiple star systems provide a complicated mix of conditions for planet formation, because the accretion potentially involves material around each star in addition to material around the group. These locations can provide opportunities as well as hazards.

Binary systems can have circumprimary (around the more massive star), circumsecondary (around the less massive star), and circumbinary (around both stars) disks, compared to likely routine planet formation sites around single stars. In widely spaced binaries you could even have protoplanetary discs around the two.

Three-dimensional numerical simulation image of young binary star system. (© Hosei University)
Three-dimensional numerical simulation image of young binary star system. (© Hosei University)

If there are no tidal effects, no perturbation from other forces, and no transfer of mass from one star to the other, a binary system is stable, and both stars will trace out an elliptical orbit around the center of mass of the system.

A multiple star system is more complex than a binary and may exhibit chaotic behavior. Many configurations of small groups of stars are found to be unstable, as eventually one star will approach another closely and be accelerated so much that it will escape from the system. This instability can be avoided if the system is hierarchical. In a hierarchical system, the stars in the system can be divided into two smaller groups, each of which traverses a larger orbit around the system’s center of mass. Each of these smaller groups must also be hierarchical, which means that they must be divided into smaller subgroups which themselves are hierarchical, and so on.

For a certain range of stellar separations, the presence of a companion star will clearly impact the formation, structure, and evolution of circumstellar disks and any potential planet formation. Global properties such as initial molecular cloud angular momentum, stellar density, the presence of ionizing sources and/or high mass, and so on, may all influence disk and thereby planet formation.

The maximum separation of bound systems is related to the stellar density. The denser clusters, in which most stars form, contain a lower fraction of bound multiple systems, comparable to the fraction found among field stars.

The binary star systems that host planets are very diverse in their properties and binary binary semimajor axes ranging from 20 AU to 6400 AU. In case where orbits are eccentric, the binary periastron can be as small as 12 AU, and important dynamical effects are expected to have occurred during and after planet formation.

In a circumbinary disk strong tidal interactions between the binary and disk are almost always expected, significantly affecting planet formation.

In a circumstellar disk with separations of a few to several tens of AU, the tidal torques of the companion star generate strong spiral shocks, and angular momentum is transferred to the binary orbit. This in turn leads to disk truncation, determining a “planet-free” zone (at least for formation). Subsequent dynamical evolution in multiple systems could still bring planets into this region.

For a circumstellar disk in a binary system, which is not influenced by strong tidal forcing, the effect of the companion star will be modest, unless the orbital inclinations are such that the Kozai effect becomes important.

Math Box 1 – The Truncation Radius

The truncation radius rt of the disk depends on the binary semimajor axis ab, its eccentricity eb, the mass ratio q = M2/M1 (M1, M2 denote the masses of the primary and secondary stars, respectively), and the viscosity v of the disk. For typical values of q = 0.5, eb = 0.3 and disk Reynold’s number of 10^5, the disk will be truncated to a radius of rt = 1/3ab.

For a given mass ratio q and semimajor axis ab an increase in eb will reduce the size of the disk while a large v will increase the disk’s radius. Not only will the disk be truncated, but the overall structure and density stratification may be modified by the binary companion.

In a circumbinary disk, the binary creates a tidally-induced inner cavity. For typical disk and binary parameters (e.g., eb = 0.3, q = 0.5) the size of the cavity is = 2.7 * ab.

Binary star diagram. Image courtesy of NASA
Binary star diagram. Image courtesy of NASA

Numerical studies of the final stages of terrestrial planet formation in rather close binaries with separations of only 20–30 AU, that involve giant impacts between lunar-mass planetary embryos, show that terrestrial planet formation in such systems is possible, if there was a possibility for planetary embryos to form.

Systems with higher eccentricity or lower binary separation are more critical for planetesimal accretion. The effects of such eccentric companion include planetesimal breakage and fragmentation because of the increased relative velocities; the circumprimary planet forming disc truncation to smaller radii, causing the removal of material that may be used in the formation of terrestrial planets; destabilization of the regions where the building blocks for these objects may exist.

For binaries with separation less than 40 AU, only very low eccentricities allow planetesimal accretion to proceed as in the standard single-star case. On the contrary, only relatively high eccentricities (at least 0.2 in the closest 10AU separation and at least 0.7 for star system semimajor at 40AU) lead to a complete stop of planetesimal accretion.

A binary companion at 10 AU limits the number of terrestrial planets and the extent of the terrestrial planet region around one member of a binary star system.

Larger periastra (> 20AU) in solar-type binary star systems with terrestrial planets formation allow the stability of Jovian planets near 5 AU. These binary star/giant planets systems effectively support volatile delivery to the inner terrestrial region.

Approximately 40–50% of binaries are wide enough to support both the formation and the long-term stability of Earth-like planets in orbits around one of the stars. Approximately 10% of main sequence binaries are close enough to allow the formation and long-term stability of terrestrial planets in circumbinary orbits. According to this, a large number of systems can be habitable, given that the galaxy contains more than 100 billion star systems, and that roughly half remain viable for the formation and maintenance of Earth-like planets.

Math Box 2 –
Stability of the satellite-type orbit, where the planet moves around one stellar component (S-Type Orbits).

In this equation, ac, the critical semimajor axis, is the upper limit of the semimajor axis of a stable S-type orbit, ab and eb are the semimajor axis and eccentricity of the binary, and mu = M2/(M1+M2). S-type orbits in binaries with larger secondary stars on high eccentricities are less stable. The +- signs define a lower and an upper value for the critical semimajor axis which correspond to a transitional region that consists of a mix of stable and unstable orbits.

Stability of the planet-type orbit, where the planet surrounds both stars in a distant orbit (P-Type Orbits).

For circular binaries, this distance is approximately twice the separation of the binary, and for eccentric binaries (with eccentricities up to 0.7) the stable region extends to four time the binary separation. A critical semimajor axis below which the orbit of the planet will be unstable is given by

Similar to S-type orbits, the +- signs define a lower and an upper value for the critical semimajor axis ac, and set a transitional region that consists of a mix of stable and unstable orbits.

Habitable zones in binary star systems

HZs in binaries depend the binaries’ orbital elements and the actual amount of radiation arriving at an orbiting planet. The analytical estimate on the extent of the HZs includes the radiation field of the binary as a function of spectral types, orbital parameters, as well as the relative orbital phase and calculations of the RMS (root-mean-square) and Min-Max distances of the inner and outer borders of the habzones in P-Type and S-Type configurations.

In a binary-planetary system, the presence of the giant planet enhances destabilizing effect of the secondary star. The Jovian planet perturbs the motion of embryos and strengthens their radial mixing and the rate of their collisions by transferring angular momentum from the secondary star to these objects.

Systems with close-in giant planets may require massive protoplanetary disks to ensure that while planetesimals and protoplanets are scattered as giant planets migrate, terrestrial bodies can form and be stable. Systems with multiple giants also present a great challenge to terrestrial planet formation since the orbital architectures of such systems may limit the regions of the stability of smaller objects.

Four different types of orbits are possible for a terrestrial planet in a binary system that hosts a Jovian planet: the terrestrial planet is inside the orbit of the giant planet; the terrestrial planet is outside the orbit of the giant planet; the terrestrial planet is a Trojan of the primary (or secondary) or the giant planet; the terrestrial planet is a satellite of the giant planet.

When numerically studying the dynamics of a terrestrial planet in a binary planetary system, integrations have to be carried out for a vast parameter-space. These parameters include the eccentricities, semimajor axes, and inclinations of the binary and the two planets, the mass-ratio of the binary, and the ratio of the mass of the giant planet to that of its host star. The angular variables of the orbits of the two planets also add to these parameters.

Except for a few special cases, the complexities of these systems do not allow analytical solutions of their dynamics, and require extensive numerical integrations. Those special cases are: binaries with semimajor axes larger then 100 AU in which the secondary star is so far away from the planet-hosting star that its perturbative effect can be neglected; binaries in which the giant planet has an orbit with a very small eccentricity (almost circular); binaries in which, compared to the masses of the other bodies, the mass of the terrestrial planet is negligible.

Instability is not the only hazard in multiple systems. The difference in masses and lifetimes can pose serious problems for life, especially in relatively close binaries or multiples.

Binary or multiple systems might be hosts to several generations of planets; life might arise and be wiped out several times in system’s lifetime. Here’s how such binary system might evolve: while both stars are on the main sequence and in close proximity to each other, small and close-in first generation of planets forms; eventually one star evolves from the main sequence into the red giant and the two stars spread further apart while stellar material blown off from the red giant builds a protoplanetary disk around the other star and second generation planets form; the second star eventually goes red giant giving the first star, which is now white dwarf, a protoplanetary disk which could create a third generation of planets.

Each generation of planets is built from stellar material with a sequentially increasing metallicity as the material is recycled within each star’s fusion processes. In this case it becomes possible for old stars, even those which formed as low metal binaries, to develop rocky planets later in their lifetimes.

However, not always changing environment might be a threat, like in case of this old gas giant PSR B1620-26 b.

Jovian planet in globular cluster M4. Credit: NASA
Jovian planet in globular cluster M4. Credit: NASA

If such planet hosted habitable satellites, and host stars remained warm and safe enough to support life, inhabitants might not been affected much by dramatic changes. Or maybe they would. But that is another story.

# Pretty much everything in this article is presently under active research. To learn more about multiple star systems and habitable planets in them, try Planets in Binary Star Systems by Nader Haghighipour, 2010; or Multiple Stars across the H-R Diagram (you can read it online), 2005.

## Have fun and more exciting stuff to follow.