Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

Architect of Worlds

by Jon F. Zeigler

Special Cases in Worldbuilding


Interim Partial Draft
Version 0.1 (31 March 2021)
Table of Contents
Special Cases in Worldbuilding ..................................................................................................................... 3
Generating Stars in Unusual Regions ........................................................................................................ 3
Open Clusters, Stellar Associations, and Moving Groups ..................................................................... 3
Inter-Arm Space .................................................................................................................................... 5
Galactic Halo ......................................................................................................................................... 6
Unusual Stars ............................................................................................................................................ 6
Massive Main-Sequence Stars .............................................................................................................. 6
Neutron Stars ........................................................................................................................................ 6
Black Holes ............................................................................................................................................ 7
Flare Stars.............................................................................................................................................. 7
Planetary Systems for Non-Main Sequence Stars .................................................................................... 8
Brown Dwarfs........................................................................................................................................ 8
Subgiant and Giant Stars ....................................................................................................................... 8
White Dwarfs and Stellar Remnants ..................................................................................................... 9
Planetary Systems for Multiple Stars ........................................................................................................ 9
Blackbody Temperatures for Multiple Radiant Sources ....................................................................... 9
Planetary Systems for Double Stars .................................................................................................... 10
Special Features for Planetary Systems .................................................................................................. 11
Asteroids and Planetoids .................................................................................................................... 11
Comets ................................................................................................................................................ 12
Kuiper Belt........................................................................................................................................... 12
Oort Cloud ........................................................................................................................................... 13
Rogue Planets...................................................................................................................................... 13
Trojan Planets ..................................................................................................................................... 14
Unusual Worlds ....................................................................................................................................... 15
Fine-Tuning World Climate ..................................................................................................................... 15
Alien Skies ............................................................................................................................................... 15
How Large Do Things Appear .............................................................................................................. 15
How Bright Do Things Appear ............................................................................................................. 16
The Color of the Sky ............................................................................................................................ 17
Unlikely Worlds ................................................................................................................................... 18
World History .......................................................................................................................................... 21
Special Cases in Worldbuilding
The world-design sequence presented in Chapter XX of this book is intended as a norm, which should
produce plausible results for most cases in an interstellar-fiction universe. However, there are special
situations in which you may wish to modify the default sequence or introduce additional detail. This
chapter covers a variety of these special cases.

Generating Stars in Unusual Regions


The default guidelines for generating star systems to place on an interstellar map (Steps One through
Eight of the design sequence, pp. XX-XX) assume a stellar neighborhood like our own. That is, a
neighborhood inside a major spiral-arm structure of the galaxy, lacking any major star-forming regions
or star clusters, and exhibiting a well-mixed population of stars of varying ages. Many regions of the
galaxy don’t fit those criteria, so star-system generation in those regions will require different
procedures.

Open Clusters, Stellar Associations, and Moving Groups


Stars form in molecular clouds or stellar nurseries of varying sizes. When a large molecular cloud
collapses, it can give rise to many young stars. When the largest and brightest of these new stars ignite,
they create radiation pressure which disperses the remnants of the molecular cloud. What remains
behind is a new open cluster, a tight grouping of stars that are all about the same age.

Open clusters are denser than a typical stellar neighborhood, with many more stars per unit volume,
especially toward their center of mass. However, they are usually not tightly bound by gravity. This leads
most of them to evaporate as they move and interact with other stars, molecular clouds, and the
galaxy’s overall structure. Over time, outlying stars are pulled away; they continue to keep a similar
velocity, following a similar orbit around the galactic core, but they slowly merge into the general
population of the galaxy.

As an open cluster evaporates, it may become a stellar association, a population of stars that are no
longer gravitationally bound and are now spread over a wide region of space, but which have a common
age and point of origin.

An older stellar association may become a moving group (also called a kinematic group). The difference
between an association and a moving group has to do with its dispersal; a moving group is usually so
broadly dispersed that astronomers may have a difficult time identifying which stars are its members.
Even so, the members of a moving group still have a common age and point of origin, and they can be
discovered by examining their chemical composition and motion through the galaxy.

It should be noted that a few open clusters appear to resist evaporation, even on a timescale of billions
of years. If an open cluster is more compact than usual, it may be gravitationally bound tightly enough to
avoid losing most of its members. A few open clusters appear to be almost as old as the galactic disk
itself (up to about 8-9 billion years old).
Placing Clusters
Within 100 parsecs of Sol, there appear to be two open clusters and as many as twelve stellar
associations or moving groups. That suggests about one such object for every 300,000 or so cubic
parsecs.

When designing an interstellar region that includes an open cluster, stellar association, or moving group,
we suggest placing a “background” population of stars using the usual random procedures, as described
in Chapters XX and XX. Then place the members of the cluster, association, or moving group in addition
to the other stars already placed. Note that it’s entirely possible for members of the galaxy’s general
population to move through the space occupied by a cluster. The result will be a region inhabited by
many stars about the same age, interspersed with outliers.

Designing Clusters
To design an open cluster or related
Cluster Age Table
object for placement on an interstellar
Tightly Bound Loosely Bound
map, start by rolling 3d6. On a 5 or less, Base Age Age Range
Roll (d%) Roll (d%)
the object will be tightly bound. 01-02 01-21 0.0 0.1
Otherwise (on a 7 or higher) it will be 03-04 22-38 0.1 0.1
loosely bound. 05-06 39-52 0.2 0.1
Next, roll 2d6, divide by 2, and make a 07-08 53-64 0.3 0.1
note of the result. This is the initial 09-10 65-73 0.4 0.1
11-12 74-81 0.5 0.1
radius of the object as an open cluster,
13-14 82-87 0.6 0.1
in parsecs. Feel free to vary the initial
15-16 88-92 0.7 0.1
radius by up to 0.25 parsecs in either
17-18 93-96 0.8 0.1
direction.
19-20 97-00 0.9 0.1
Next, roll 2d6 once more, with a 21-45 - 1.0 2.0
minimum of 7 if the cluster is tightly 46-00 - 3.0 5.0
bound. Divide the result by 2, then multiply by the cube of the cluster’s initial radius in parsecs. This will
generate an estimate for how many star systems are members of the original cluster before any
evaporation takes place.

Now roll d% on the Cluster Age Table, using either the Tightly Bound or the Loosely Bound column as
appropriate. To determine the cluster’s exact age, roll d% again, treat the result as a number between 0
and 1, multiply that number by the Age Range, and add the result to the Base Age. The result will be the
cluster’s age (and therefore the age of each star system in the cluster) in billions of years. You may wish
to round the age to two significant figures.

When placing stars of the cluster on the interstellar map, begin by selecting a specific point in space as
the center of the cluster. The stars of the cluster should be placed in three zones:

• The cluster core zone is within the cluster’s initial radius from the center point.
• The tidal radius zone is from 1-4 times the cluster’s initial radius from the center point.
• The extended halo zone is from 4-20 times the cluster’s initial radius from the center point.

To estimate how many stars to place in each of these zones, refer to the Cluster Evaporation Table:
Cluster Evaporation Table
Cluster Age Core Tidal Radius Extended Halo
0 100% 0% 0%
0.1 80% 20% 0%
0.2 64% 32% 4%
0.3 51% 38% 10%
0.4 41% 41% 15%
0.5 33% 41% 20%
0.6 26% 39% 25%
0.7 21% 37% 28%
0.8 17% 33% 29%
0.9 13% 30% 30%
1.0 or more 11% 27% 30%

Here, the Cluster Age is in billions of years, and the three percentages indicate roughly how many of the
cluster’s initial star systems should be placed in each zone. Note that from about 300 million years on,
the percentages do not add up to 100%. This reflects that some cluster members “evaporate” entirely
and are lost among the general population of the galaxy.

Tightly bound clusters “evaporate” much more slowly. For such clusters, divide the cluster’s age by 10
before referring to the table.

Inter-Arm Space
It’s a common misconception that the space between the galaxy’s spiral arms is empty, lacking in stars
and other objects of interest. This is far from being the case.

The spiral arms are certainly regions of greater density within the interstellar medium, the matter that
exists in the space between star systems throughout the galaxy. This greater density tends to cause the
interstellar medium to collect, forming the great molecular clouds that eventually give rise to new stars.
As families of young stars appear, they tend to illuminate the molecular clouds in which they formed,
often making them visible over very long distances. Meanwhile, those few young stars which are very
massive are also extremely bright, shedding as much light as tens of thousands to millions of their
smaller siblings. This combination of illuminated nebulae and super-bright young stars is what makes the
spiral arms so clearly visible, even in distant galaxies.

Molecular clouds and super-bright stars rarely last more than a few million years, so after their
formation they don’t have time to move out of the spiral arm where they formed. On the other hand,
small, more long-lived stars have plenty of time to leave their original spiral arm, traversing inter-arm
space and other spiral arms many times as they orbit the galaxy’s core. Our own sun has circled the
galaxy about twenty times since its formation, and it has crossed spiral-arm structures many times. In
fact, it’s something of an accident that we find ourselves inside a spiral arm (more precisely, the Orion
Spur, which connects two major spiral arms) at this specific point in time.

When generating a map for a region of inter-arm space, it might be reasonable to reduce the density of
star systems, but not to a marked degree. Consider placing at least 80% as many star systems per
volume as the default discussed in Chapter XX.
The major difference will be in the age of any star systems present. Assume stars in inter-arm space are
at least 50 million years old – if any star system is generated in Step Four with age less than this, re-roll
the system’s age. Open clusters and stellar associations appear in inter-arm space as well, but again they
are likely to be at least 50 million years old.

Galactic Halo
TO BE COMPLETED

Unusual Stars
By far, most objects on an interstellar map will be those that can be generated using the standard design
sequence: brown dwarfs, main-sequence stars between 0.08 and 6 solar masses, subgiant or giant stars
at the end of their lives, and the stellar remnants left behind by such stars. In the real universe there is
even more variety, unusual objects that we might choose to place as a rare case. In this section we will
discuss several options.

Massive Main-Sequence Stars


The frequency of stars at a given mass appears to follow a power law, so that more massive stars are
less and less common. The most massive stars known have up to about 200 solar masses; there is some
theoretical work to suggest that this is a loose upper limit for stability.

Stars above 6 solar masses are extremely rare; perhaps one out of a thousand stars form with that much
mass, and such stars have very short lifespans so that it’s even more unusual to see one still burning.
The nearest such main-sequence star appears to be Achernar (Alpha Eridani), about 140 parsecs from
Earth.

You may wish to place a massive star on an interstellar map as a rare case. Once you have determined
the star’s original mass, use the following to approximate its other parameters:

𝐿 = 𝑀4

𝑆 = 10 × 𝑀−3
Here, M is the star’s mass, L is its approximate luminosity, and S is an estimate for its main-sequence
lifespan in billions of years. Stars with this level of mass will be of spectral types B or O during their main
sequence lifespan, and they will expand to become supergiant stars late in their lives.

Note that such massive stars have very short stable lifespans. Unless the star is very young, it will likely
have already passed through its subgiant and giant phases and become a stellar remnant. Such stars are
unlikely to form planetary systems, since they expire before the process of planetary formation can
advance. You may wish to restrict placement of such stars to very young open clusters (p. XX).

Neutron Stars
If a main-sequence star begins life with roughly 10-25 solar masses, it will be too massive to form a
white dwarf at the end of its life. Instead, the remnant left behind by its supernova explosion will be a
neutron star, a mass of neutrons and more exotic particles only a few kilometers across.

Neutron stars will normally have mass between about 1.4 solar masses (the Chandrasekhar limit) and
about 2.1 solar masses (the Tolman-Oppenheimer-Volkoff limit). It’s possible that some neutron stars
will exceed this upper limit, depending on their internal structure. The theoretical upper limit for a
neutron star is about 2.7 solar masses; any stellar remnant more massive than this is almost certainly a
black hole (see below).

Neutron stars tend to be very hot, with surface temperatures up to 600000 K. They also rotate very
quickly – some very young neutron stars have been observed to rotate in a tiny fraction of a second.
Despite their extremely high effective temperature, neutron stars are so small that they have negligible
luminosity.

Neutron stars often have very powerful magnetic fields, causing most of their radiation to emerge from
near their magnetic poles. If the magnetic poles aren’t lined up with the neutron star’s rotational axis,
then these emission beams will sweep the sky with a very predictable pattern. Such objects are also
called pulsars.

Neutron stars are very rare and should not be placed on an interstellar map at random. For comparison,
the known neutron stars nearest to Earth are on the order of 120-150 parsecs away.

Black Holes
The stellar remnant of a main-sequence star which begins life with 25 or more solar masses will be too
massive to stabilize as either a white dwarf or a neutron star. Instead, the star’s core will become a black
hole.

Stellar-mass black holes are expected to have mass of at least 2.1 solar masses, although some stellar
remnants of up to about 2.7 solar masses may survive as neutron stars. The smallest candidate black
holes so far detected seem to be at least 3.0 solar masses, and most are significantly larger.

Strictly speaking, a stellar-mass black hole isn’t a material object in space. It’s a very compact region
(about 30 kilometers across) with mass, electric charge, and angular momentum. The gravity of this
region is so strong that nothing, not even light, can escape. A black hole is utterly dark, and it can only
be detected by the effect of its gravity. Matter that falls into the black hole will be annihilated, releasing
a flood of radiant energy on the way. A black hole’s gravity also bends light that passes by, causing an
effect called “gravitational lensing.”

Black holes are extremely rare and should not be placed on an interstellar map at random. The nearest
known black-hole candidate to Earth is 450-500 parsecs away. It’s possible that one or more black holes
are closer than this, of course. For example, a black hole that originally formed as a companionless star
and is currently moving alone through interstellar space might be very difficult to spot.

Flare Stars
Almost all main-sequence stars exhibit flare activity. That is, they experience intense outbreaks of
electromagnetic radiation in their atmospheres, usually ultraviolet or X-ray light. These flares are often
accompanied by coronal mass ejections, along with eruptions of the fast-moving charged particles that
make up the stellar wind. Flares are caused by the release of energy stored in magnetic fields as they
interact with the star’s atmosphere.

Stellar flares are an issue for any worlds in a star’s planetary system. Coronal mass ejections and stellar
wind associated with flares can interact with a planet’s atmosphere, stripping atmospheric gases away
to space. Meanwhile, if the UV and X-rays from a flare are not filtered out by a world’s atmosphere, they
will bombard the surface and pose a threat to any living things there.
Although most stars flare, the phenomenon is particularly marked for small, cool stars. Some red dwarfs
produce frequent and intense flares, like or even exceeding those of a much larger and brighter star. The
result is a dramatic increase in the star’s luminosity that can last several minutes. Much of the flare’s
light will be in the short-wavelength bands that the star normally doesn’t produce. Worlds orbiting such
a flare star, probably already very close to their primary, will be subjected to the full force of its flares.

The effect of a flare star’s eruptions on a world’s atmospheric retention is already “baked into” the
standard design sequence. Worlds with strong magnetic fields should be able to keep all but the most
powerful outbursts of stellar wind at bay, preserving their atmospheres over long timescales.

On the other hand, you may wish to note whether a red dwarf is likely to be a flare star, producing
frequent, intense flares. Such a primary star will be particularly hostile to exposed life on its planets. To
estimate whether a red dwarf star is likely to flare frequently, roll 3d6, modified as follows:

• +1 for every full billion years that the star’s age is less than 6 billion years
• +1 for spectral types M4 or M5
• +2 for spectral types M6 or M7
• +3 for spectral type M8

The red dwarf will be a flare star on a result of 15 or higher. Remember that all main-sequence stars
flare, and even older or brighter red dwarfs will still produce occasional large flares. A “flare star” is one
for which the flares are frequent and unusually intense.

Planetary Systems for Non-Main Sequence Stars


The main design sequence laid out in Chapter XX should work without modification for stars on the main
sequence, including red dwarf stars. To generate planetary systems for objects outside this range, some
of the sequence (specifically Steps Nine and onward) may need to be adjusted.

Brown Dwarfs
When designing a planetary system for a brown dwarf star, assume that the brown dwarf’s initial
luminosity was about 0.00075 solar units. Naturally, the current conditions on any of its planets will
depend upon its current luminosity, which will be lower rather than higher.

Note that brown dwarfs old enough to have life-bearing planets will almost certainly have cooled to the
point that they give off no significant visible light. If a world’s brown dwarf primary is of spectral L, T, or
Y, then photosynthesis will not be possible. Take note of this when determining the world’s biological
history in Step Twenty-Six.

Subgiant and Giant Stars


Planetary systems for subgiant and giant stars can be designed, for the most part, using the normal
sequence. However, for post-main-sequence stars it’s entirely possible for one or more inner planets to
be destroyed by the primary’s expansion.

If any planet’s minimum distance from its primary star (as computed in Step Twelve) is less than the
star’s current radius, eliminate the planet from the system (the star has been absorbed by its expanding
primary). Also, be sure to check the case in Step Nineteen regarding blackbody temperature so high as
to indicate the vaporization of rock.
White Dwarfs and Stellar Remnants
Planetary systems can survive the red-giant phase of a star’s lifespan, remaining in place after the star
has become a white dwarf or a more compact stellar remnant. Astronomers have detected exoplanets
orbiting white dwarfs and even neutron stars. In fact, some such exoplanets appear remarkably close to
their primaries, as if they somehow survived the red-giant phase, or they migrated inward afterward.
The mechanisms by which this might happen are not yet well understood, so designing a plausible
planetary system for a stellar remnant involves a certain amount of guesswork.

To design such a surviving planetary system, work from the star’s original mass and luminosity, as if it
was still on the main sequence. Once the design process has reached the end of Step Twelve, eliminate
any planets whose minimum distance from the primary star is less than 1.0 AU. These planets have been
engulfed and destroyed by the star during its red-giant phase. Also eliminate any Planetoid Belt from the
original planetary system.

Next, modify the orbital radii of surviving planets to account for migration after the red-giant phase:

• Begin with the innermost surviving planet. To determine its final orbital radius, make another
unmodified roll on the Planetary Migration Table on p. XX and multiply the planet’s original
orbital radius by the resulting factor. In an extreme case (a Migration Factor of 0) the planet’s
final orbital radius will be 1d6+1 times 0.01 AU.
• Next, place the outermost surviving planet. To determine its final orbital radius, roll 3d6+4 on
the Planetary Migration Table; the Migration Factor for this planet is equal to the result, or to
the result of the first roll, whichever is greater. Multiply the outermost surviving planet’s original
orbital radius by the result.
• Finally, redistribute any remaining surviving planets between the innermost and outermost,
using the procedure on p. XX in Step Ten.

At this point, the surviving worlds of the planetary system can be developed according to the standard
procedures in Steps Thirteen and onward.

Planetary Systems for Multiple Stars


The design sequence in Chapter XX generally assumes that worlds will have only one primary star, to
serve as the gravitational center and source of radiant energy for a planetary system. This assumption
becomes problematic when considering worlds in a multiple star system. For example, it’s possible for a
world to orbit one star in a gravitationally bound pair, while the other star of the pair is close enough to
provide a significant amount of radiant energy, raising the world’s temperature. It’s also possible for a
world to orbit both stars of a bound pair (in this case, the world technically orbits the barycenter or
“center of mass” of the double star). Both cases may require adjustments to the standard design
sequence.

Blackbody Temperatures for Multiple Radiant Sources


When computing the blackbody temperature for a given world (Step Nineteen on p. XX), if there are
multiple sources of radiant energy, the formula found on p. XX should be expanded to:

4 𝐿𝑘
𝑇 = 278 × √∑
𝑅𝑘2
𝑘
That is, for each brown dwarf or star in the system, divide the luminosity by the square of the average
distance. Then sum all those quantities, take the fourth root of the result, and multiply by 278 K.

Example (Alice): Alice is computing the blackbody temperature for a world that orbits a red dwarf star
with luminosity of 0.0005 solar units, at an average distance of 0.025 AU. Meanwhile, the red dwarf is
the companion of a brighter star, with luminosity of 1.0 solar units. The two stars orbit at an average
distance of 2.0 AU, and so the world’s average distance from the brighter star is also 2.0 AU.

If Alice were to take only the red dwarf into account, the blackbody temperature of the world would be:
4
√0.0005
𝑇 = 278 × ≈ 263 K
√0.025
However, Alice needs to consider the radiant energy provided by the brighter primary star as well. The
computation therefore becomes:

4 0.0005 1.0
𝑇 = 278 × √ + ≈ 281 K
0.0252 2.02

Clearly, the primary star of the system, even though it is at a much greater average distance, provides
enough energy to noticeably warm the world in question.

Example (Bob): Bob is computing the blackbody temperature for a world that orbits a pair of stars. The
two stars have luminosity of 0.8 and 0.6 solar units respectively, and they orbit one another at an
average distance of 0.3 AU. The world orbits the barycenter of this system at an average distance of 1.5
AU, so Bob uses this figure for the average distance to each star.

The computation works out as:

4 0.8 0.6
𝑇 = 278 × √ 2 + ≈ 247 K
1.5 1.52

Planetary Systems for Double Stars


When designing planetary systems for multiple stars, planets can form under two different cases:

• A planet can orbit one member of the multiple star system (this is called an S-type orbit)
• A planet can orbit both members of a gravitationally bound pair (this is called a P-type orbit, and
the planet is sometimes called a circumbinary planet)

Planetary Systems in S-Type Orbits


Planets which form in S-type orbits are already covered in the standard design sequence.

Note that the planetary system which forms around a single component of a multiple star will have a
forbidden zone (p. XX) which marks the band of possible orbits that will not be stable in the long term.
Inside the forbidden zone’s inner radius, the gravitational influence of the other stars can be assumed to
be negligible.

Technically, the presence of stellar companions will tend to warm a protoplanetary disk, suggesting that
the formation orbits (p. XX) should be placed further out than in the single-star case. Since the relevant
computations are complex and the formation-orbits model is only an approximation in the first place,
we suggest ignoring this factor. Assume that the protoplanetary disk in an S-type case is affected only by
its central star.

Planetary Systems in P-Type Orbits


In some cases, you may wish to design a planetary system for the P-type case. This is especially likely to
be worthwhile if the pair of stars is at Extremely Close or Very Close separation (that is, the average
distance between them is no greater than about 1.5 AU).

In this case, the planetary system can be designed with minimal modification to the standard design
sequence. Carry out all computations in Step Nine and onward as if the binary pair was a single star
located at the barycenter of the system. This notional star’s mass should be the sum of the masses of
both components. Likewise, its luminosity should be the sum of the luminosities of both components;
use this computation both for the initial luminosity when laying out the protoplanetary disk, and for the
current luminosity when determining each world’s characteristics.

Note, however, that planetary orbits too close to the barycenter will not be stable. The binary pair will
create an inner forbidden zone. To take account of this, multiply the maximum separation of the two
stars by 3.5, and treat the result as the disk inner edge radius for all purposes in Steps Ten and Eleven.

Special Features for Planetary Systems


Aside from the “worlds” (planets and their moons) generated by the design sequence, planetary systems
will include a variety of other objects which might be of interest in a story or game.

Asteroids and Planetoids


Asteroids, also called planetoids, are small stony bodies. The standard design sequence often places
planetoid belts, usually just inward from a system’s dominant gas giant. In a typical belt, the largest
objects will usually be about 250-500 kilometers in radius. A few hundred planetoids may be at least 25-
50 kilometers in radius, and there may be millions of planetoids about 1 kilometer in radius. Despite
these huge numbers, the total mass of a planetoid belt will not add up to the mass of even a small
planet. Planetoid belts are also not the dense “asteroid thickets” found in some science-fiction media.
Spaceships will be able to cross through them without significant danger, and it will be a rare occasion to
see even one with the naked eye.

In a planetoid belt, the objects are not all found at the same orbital radius – the radius generated in the
design sequence should be considered an average. Planetoid orbits are often mildly eccentric or inclined
to the system’s ecliptic plane.

Many planetoids may be found away from the main planetoid belt of a system, or even in systems that
have no prominent planetoid belt. For example, the trojan positions in the orbit of a dominant gas giant
(see p. xx) are likely to collect many planetoids. Other planetoids may fall into orbital resonance with the
system’s planets, becoming “quasi-satellites” for even relatively small terrestrial worlds. Still others may
be captured, becoming a planet’s moonlets (p. XX).
Planetoids are classified according to their composition:

• C-type planetoids, also called carbonaceous planetoids, are very dark in color, composed of light
compounds and minerals. They are likely to have carbon materials, and possibly some water in
the form of hydrated compounds or water ice. C-type planetoids tend to be very old, left over
from the formation era of the planetary system. They are quite common, usually amounting to
about 75% of all planetoids.
• S-type or stony-iron objects are composed mostly of silicates and stony minerals, with a
significant amount of metal. They make up about 15% to 20% of all planetoids.
• M-type or metallic objects are the rarest, making up 5% to 10% of all planetoids. They are
composed almost entirely of metal, mostly nickel and iron with traces of other metals. These are
likely to be popular among asteroid miners and space-faring civilizations.

Comets
Comets and other icy bodies are normally found in the cold outer reaches of a planetary system. Most
comets are between 1 and 50 kilometers in diameter. They are often described as “dirty snowballs” –
their composition is dominated by water and other ices, but they also include traces of silicate dust,
metals, and organic compounds. Any given star system may have trillions of comets, their mass adding
up to that of a small planet.

Comets are sometimes redirected into the inner planetary system, usually due to gravitational
interaction with some larger object. When this happens, the comet takes up an eccentric orbit,
sometimes approaching the primary star so closely that its ices boil off to form a cometary “tail.” After
passing through the inner planetary system several times, a comet may meet with a variety of fates. It
may fragment after losing most of its icy materials, adding to the inner system’s inventory of asteroid
debris. It’s also possible that a close encounter with a planet may further modify its orbit, causing it to
take up a planetoid-like path that remains in the inner planetary system. A comet can also ends its
career by colliding with a planet!

Kuiper Belt
Many comets exist in a flattened disk, relatively close to the primary star, called the Kuiper Belt. This
structure appears during planetary formation in formation orbits 10-12, especially if this region falls
outside the protoplanetary disk’s slow-accretion line. In effect, many icy planetesimals form in that
region, but because they move so slowly in their orbits, they have no opportunity to coalesce into a full-
fledged planet before the formation era comes to an end. The Kuiper Belt is the primary source for so-
called “short-period comets,” those whose regular visits to the inner planetary system are only years or
decades apart.

The inner edge of the Kuiper Belt begins at the system’s slow-accretion line, and may extend outward
for many AU. Aside from billions of comets, the belt may include several larger dwarf planets, round icy
bodies with up to about 0.002 to 0.003 Earth-masses.

The Kuiper Belt in our own planetary system appears to have been depleted billions of years ago,
possibly when a Nice Event (p. XX) pushed the gas giant Neptune out into what had once been the heart
of the belt. Astronomers have uncovered evidence that other star systems have much denser Kuiper
Belt structures. For example, the Tau Ceti system’s Kuiper Belt may be as much as ten times as dense as
our own.
To estimate the relative density of a system’s Kuiper Belt, begin by considering how much total
planetesimal mass is placed in formation orbits that fall outside the system’s slow-accretion line, but not
in a forbidden zone. (Naturally, a star that has a forbidden zone may not have a Kuiper Belt at all.)

Then, if any Gas Giant planet ends up outside the slow-accretion line due to a Nice Event, divide this
total planetesimal mass by four. Compare the result to the value for our own planetary system (about 2)
to get a sense for whether the Kuiper Belt is thinner or denser than our own, and by how much.

Oort Cloud
Far beyond a planetary system’s outer edge, a star system will likely possess an Oort cloud. This is a
reservoir of comets in nearby interstellar space, stretching from several hundred AU out to as far as 1
parsec. Unlike the Kuiper Belt, the Oort cloud will form an even sphere around the star system. The Oort
cloud forms a reservoir for long-period comets, which can approach the inner system from any direction
and usually have orbital periods of thousands or even millions of years.

The Oort cloud may also contain larger objects, such as the icy dwarf planets also found in the Kuiper
Belt, or even full-sized planets that were ejected from the planetary system during the formation era.
Such objects may be very difficult to detect; so far, astronomers have found none in our own Oort cloud.

Rogue Planets
During the formation of a planetary system, it’s possible for one or more planets to be ejected to
interstellar space. It’s also possible for planets to form in a nursery cloud, independent of any primary
star. Astronomers have observed dozens of these rogue planets, some with several times Jupiter’s mass,
a few as small as Earth.

Rogue planets are very difficult to detect, since they are small and dark in the depths of interstellar
space. Astronomers usually detect them by observing microlensing events, in which the mass of a rogue
planet briefly distorts light from a distant star. Estimates for the number of rogue planets in the galaxy
vary widely, but at this writing the upper limit seems to be about one Jupiter-mass rogue planet for
every four main sequence stars. Smaller rogue planets are probably more common.

Rogue planets can provide “stepping-stones” between luminous stars, which might be particularly useful
for civilizations that have no access to faster-than-light travel. Feel free to place rogue planets on your
interstellar map as needed. Here are some guidelines for applying the design sequence to these worlds.

Age: Use the procedures under Step Four (pp. XX-XX) to determine the age of a rogue planet, as if it
were an independent star system.

Mass: At present, we have no way to estimate the distribution of masses for rogue planets. Consider
that a rogue planet can be a Leftover Oligarch, a Terrestrial World, a Failed Core, or a Gas Giant of
varying size. Select any of those categories and assign an appropriate mass. Remember that a planet
with greater than 4000 Earth-masses is a brown dwarf and should be placed using the procedure for
placing stars.

Density, Radius, and Surface Gravity: Determine these parameters based on the rogue planet’s category
and mass, as usual.

Natural Satellites: Rogue planets which are Gas Giants are very likely to have satellites that form by
natural accretion. Rather than computing the planet’s Hill radius, simply roll 1d+2 and generate that
many major satellites using the procedures on pp. XX-XX. Smaller rogue planets are unlikely to retain
satellites; in this case roll d%, and on a result of 96-00 the rogue may have either a major satellite or one
or more moonlets, generated using the procedures on pp. XX-XX.

Obliquity: Since rogue planets do not orbit any primary star, there is no meaningful measure of their
obliquity. Skip this step entirely. They also have no “day” or “year” based on their movement around a
luminous star.

Blackbody Temperature: Since a rogue planet has no primary, it has no source for incoming energy.
Assume that a rogue planet (and any of its satellites) has blackbody temperature of 30 K and compute
its M-number from that figure.

Prevalence of Water: Assume that a non-Gas Giant rogue planet formed outside the formation ice line.
Most such rogue planets will have Massive prevalence of water and 100% hydrographic coverage.

Geophysical Parameters: Compute only the Primordial and Radiogenic Heat Budget for any non-Gas
Giant. Determine the world’s lithosphere and plate tectonics based solely on that figure.

Magnetic Field: Determine the strength of a rogue planet’s magnetic field normally.

Planetary Atmosphere and Native Life: Determine a non-Gas Giant rogue planet’s atmosphere normally.
Since the planet’s blackbody temperature is so low, it will almost certainly have nothing but hydrogen
and helium in the atmosphere. Note that it’s entirely possible for a rogue planet to have life originating
in deep hydrothermal vents. Photosynthesis will not be possible.

Trojan Planets
It’s possible for a small celestial body to share the orbit of a much larger one, remaining close to a stable
position 60° ahead or behind the main body in the shared orbit. Such an object is called a trojan. The
positions where a trojan can reside are two of the main body’s Lagrange points. The position ahead of
the main body is usually labeled L4, while the position behind the main body is labeled L5.

In our own planetary system, several of the planets have trojan planetoids. Venus, Earth, and Mars each
have at least one or two trojans, although these are probably only temporary trojans that will eventually
be perturbed away from their stable positions. Jupiter is the undisputed king of trojan planetoids, with
many thousands of them. In fact, there appear to be as many planetoids in the Jupiter trojan positions
as there are in the main asteroid belt.

It appears to be at least possible, if rare, for a planet to form as a trojan. There is some evidence that the
young Earth had a Mars-sized trojan called Theia. In this hypothesis, the conditions for the trojan’s long-
term stability were not met, so Theia eventually wandered away from its Lagrange point and collided
with Earth. Had Earth been more massive, its Lagrange points would have been more stable against the
perturbation of other planets, and Theia might have survived to the present day.
At present, we have yet to detect any exoplanet trojans, and we have no way to estimate how common
trojan planets might be in other planetary systems. If you would like to place a trojan planet as a rare
case, check the following guidelines:

• The main planet in the relationship should be the system’s dominant gas giant.
• The dominant gas giant should have migrated inward from its original formation orbit,
eliminating at least one inward formation orbit in the process.
• No Grand Tack should have taken place.
• The dominant gas giant’s orbit should have eccentricity no greater than 0.05.
• The primary star must have mass at least 100 times that of the dominant gas giant. Remember
that one solar mass unit is equal to about 330,000 Earth-masses.

If all the above conditions are met, then it would be at least possible for a trojan planet to be stable over
billion-year timescales. The trojan planet may be a Leftover Oligarch, a Terrestrial Planet, or a Failed
Core. Its mass may be no greater than 1% that of the dominant gas giant, and it is likely to be
significantly less than that.

Unusual Worlds
TO BE COMPLETED

Fine-Tuning World Climate


TO BE COMPLETED

Alien Skies
Anyone developing a setting for interstellar fiction is likely to be interested in describing how objects
appear in the sky of an alien world. How large and how bright are that planet’s moons? What stars are
brightest in the sky? Do any of the other worlds in the system appear in the sky? What kind of colors can
be found on alien worlds? This section provides guidelines for developing those descriptions.

How Large Do Things Appear


To determine the angular diameter of an object in the sky, use the following:
𝑑
𝛿 = 2 sin−1 ( )
2𝐷
Here, δ is the angular diameter of the object in degrees, d is the object’s diameter, and D is the distance
to the object. Note that d and D need to be measured in the same units of distance, but it doesn’t
matter what units those are – the formula will work regardless.

For comparison, consider the following:

• An adult human’s fist held at arm’s length is about 10° across.


• An adult human’s three middle fingers, held at arm’s length, are about 5° across.
• An adult human’s little fingertip, held at arm’s length, is about 1° across.
• Both the Moon and the Sun have angular diameter of about 0.5° in Earth’s sky; any object that
appears at least this large will show clearly to the human eye as a small disc.
• The unaided, normal human eye can resolve objects down to about 1 minute of arc, or one
sixtieth of a degree. Anything smaller than this will appear to be a starlike point.
How Bright Do Things Appear
Astronomers use a magnitude scale to measure
Bolometric Correction Table
the brightness of celestial objects. Magnitude is
Spectral Spectral Spectral
defined as an inverse logarithmic scale: lower B B B
Type Type Type
values indicate brighter objects, and if one object A0 0.30 F6 0.14 K3 0.52
has magnitude exactly 5 less than another, that A1 0.25 F7 0.15 K4 0.62
means it is 100 times as bright. By computing an A2 0.20 F8 0.16 K5 0.72
object’s apparent magnitude in the V-band (for A3 0.19 F9 0.17 K6 0.85
“visual light”), we can gain some insight in how A4 0.17 G0 0.18 K7 0.98
bright it will appear in the sky of an alien world. A5 0.15 G1 0.19 K8 1.11
A6 0.14 G2 0.20 K9 1.24
Absolute Visual Magnitude for Stars
A7 0.13 G3 0.20 M0 1.38
To compute the absolute visual magnitude for a
A8 0.12 G4 0.21 M1 1.63
star, use:
A9 0.11 G5 0.21 M2 1.89
𝑀𝑉 = −2.5 log10 𝐿 + 4.74 + 𝐵 F0 0.09 G6 0.23 M3 2.17
F1 0.10 G7 0.25 M4 2.45
Here, MV is the absolute visual magnitude for the F2 0.11 G8 0.27 M5 2.73
star, L is the star’s luminosity, and B is the F3 0.12 G9 0.29 M6 3.01
bolometric correction for the star’s spectral type. F4 0.13 K0 0.31 M7 3.29
Refer to the Bolometric Correction Table to the F5 0.14 K1 0.36 M8 3.57
right to find the appropriate value for B. K2 0.42 M9 3.85
Note that the bolometric correction is necessary because a star’s energy output will not be entirely in
the visible band. The larger the value of B, the more of the star’s luminosity is in the ultraviolet or
infrared ranges. Notice that for red dwarf stars, very little of the star’s energy output may be in the form
of visible light!

Apparent Visual Magnitude for Stars


Once the absolute visual magnitude for a star is known, its apparent visual magnitude can be computed.

Compute the distance modulus μ as follows:

𝜇 = −5 + 5 log10 𝐷
Here, D is the distance to the star in parsecs. If it would be more convenient to measure the distance to
a star in AU, use the following instead:

𝜇 = −31.6 + 5 log10 𝐷
In this version of the formula, D is measured in AU.

To compute the star’s apparent visual magnitude, simply add the distance modulus to the absolute
visual magnitude.

Absolute Visual Magnitude for Worlds


Astronomers define absolute magnitude differently for planets or other planetary-system objects than
they do for stars. The usual variable for the absolute magnitude of a planet is H rather than M, and can
be computed as follows:
𝐻𝑉 = 9.37 + 𝑀𝑉 − 5 log10 𝑅 − 2.5 log10 𝐴
Here, HV is the world’s absolute magnitude, MV is the absolute magnitude of its primary star, R is its
radius in kilometers, and A is its albedo.

Apparent Visual Magnitude for Worlds


TO BE COMPLETED

The Color of the Sky


Alien worlds are very likely to have daytime skies whose color differs from that of Earth. While it’s
difficult to calculate exactly how any given world’s sky will appear, there are a few useful rules of thumb.

A world’s daytime sky takes on some color because of the scattering of light in the atmosphere. Most of
the available light comes directly from the primary star, but some of it will interact with particles in the
atmosphere, in such a way as to change their trajectory in a random direction. As a result, light comes to
the observer along every path that intersects the atmosphere, appearing to flood the entire sky with
light of varying color. If there is no atmosphere, no scattering will take place, and the sky will be a star-
spangled black whether the primary star is visible or not. However, even a very thin atmosphere – such
as would be found on a Class 5 world with no atmospheric mass, like Mars – can suspend enough fine
particulates to cause significant scattering.

Rayleigh Scattering
The dominant form of scattering in the atmosphere of most worlds is Rayleigh scattering. This involves
scattering of light by particles that are much smaller than the wavelength of the radiation, such as the
air molecules in a substantial atmosphere. When the primary star is high, Rayleigh scattering tends to
work more on the shorter wavelengths of visible light, and the result is the blue color that we find
familiar.

As a rule of thumb, if a world’s atmospheric mass is less than 1, the color due to Rayleigh scattering
becomes a deeper blue, shading to violet and then to black for very thin atmospheres. For example, the
Martian atmosphere generates almost no Rayleigh scattering, so its daytime sky would probably be
black were it not for other forms of scattering (see below).

If a world’s atmospheric mass is greater than 1, scattering will extend to longer and longer wavelengths,
so that the color of the sky pales and becomes closer to a uniform white. Worlds with atmospheric mass
of greater than 2-3 may have pearly white skies in daytime.

As the primary star approaches the horizon, its light is forced to travel through more of the atmosphere,
and through thicker layers of air close to the ground. This tends to remove some shorter-wavelength
light, causing the star and the nearby sky to appear reddened. As with the high-sun sky color, this effect
will be weakened in thinner atmospheres and strengthened in denser.

The spectral type of the primary star has less effect on the color of the sky than might be expected. Only
for primary stars of types K8, K9, M0, M1, or M2 will the color of the sky begin to noticeably change, as
the starlight loses more and more of the “blue” wavelengths. A pearly white sky in high daytime will be
more common even for Earthlike or lower atmospheric mass. Very cool red dwarf stars (M6, M7, or M8)
provide so little blue light that the skies are likely to be reddish even when the primary star is high.
Mie Scattering
Mie scattering takes place when light is scattered by particles that are larger than the wavelength of the
radiation. Even on worlds with substantial atmospheres, Mie scattering can affect the color of the sky.
On worlds with very thin or trace atmospheres, Mie scattering may be dominant.

The chemical composition of the particles causing Mie scattering can be critical in determining the
resulting color. For example, on Earth the relevant particles are usually high-altitude ice crystals, water
droplets, or sometimes air pollutants, all giving rise to a white or greyish haze. On Mars, the upper
atmosphere contains fine particles dominated by iron oxides, giving rise to a daytime sky that is
butterscotch in color. On Titan, high-altitude hydrocarbon haze causes the sky to take on a dull yellow or
yellow-brown color.

Green skies are often seen on science-fictional planets, but in reality these are unlikely to occur on
worlds otherwise habitable for humans. One possibility would be worlds on which large amounts of
microscopic photosynthetic life are suspended in the upper atmosphere. Such (“sky plankton”) is likely
to be rare but not entirely implausible; even on Earth, bacteria and other single-celled organisms can be
found in the upper air.

Unlikely Worlds
Science-fiction media often present exotic alien environments that turn out to be unlikely in the real
universe, at least if our current understanding of planetary formation and dynamics is correct. The
standard design sequence in Chapter XX avoids most of these exotica. However, unlikely doesn’t mean
impossible, so it might be reasonable to place such a situation in an interstellar setting as a rare case.

Multiple Moons
In the standard design sequence, a non-Gas Giant planet is only likely to acquire a single major satellite,
after a major impact late in the process of planetary formation.

In theoretical models of this process, material thrown into orbit by the collision will usually form only
one satellite. In cases where two or more satellites form, the situation is normally unstable, due to
gravitational interaction between the satellites or perturbations from other planets. Before long, the
“extra” satellites fall back to the source planet, collide with one another, or are ejected from the system.
Eventually, only one major satellite is left. For example, there is some evidence that Earth originally
formed two moons after the impact of Theia, but the smaller of the two eventually collided and merged
with the object that eventually became our sole Moon.

One circumstance in which multiple large satellites might survive is if they fall into an orbital resonance
very early in the process. Orbital resonances tend to stabilize the dynamics of the system, keeping the
orbiting bodies spaced out and giving them some resistance to outside perturbation. That suggests the
following procedure to modify the second case in Step Fourteen.

Begin by selecting the number of major satellites, if greater than one. By far the most likely case is
exactly two major satellites.

The total mass of all surviving major satellites will still be about one hundredth of the mass of the
planet. Determine the mass of each major satellite using the formula on p. XX, but then divide each
result by the number of major satellites you plan to place.
To determine the orbital radius of the innermost satellite, do the 3d6+7 roll on p. XX twice, then take the
lesser of the results and base the orbital radius on that.

To place the rest of the major satellites, use only those ratios between orbital radii that will produce
strong orbital resonances:

Strong Orbital Resonances Table


Ratio Resonance
1.211 4:3
1.251 7:5
1.310 3:2
1.368 8:5
1.406 5:3
1.452 7:4
1.480 9:5
1.587 2:1 (Laplace)

Note that the Laplace resonance should only be used if there are at least three major satellites, since it
is only (and exceptionally) stable in a stack of at least two ratios. In the very rare case of three or more
major satellites, the Laplace resonance might be specifically useful in ensuring the stability of the
system.

Very Large-Appearing Moons


Visual media often present enormous objects in the skies of alien worlds, satellites (or other planets)
that appear far larger than our own Moon does in Earth’s sky. This can be very evocative, but there are
several reasons why it seems unlikely.

To begin, models of the large-impact process for the formation of a major satellite seem to indicate that
the resulting object will be small in comparison to its parent. Not that much of the mass of the planet (or
its impactor) is thrown into space in such a way as to be available for formation of the satellite. For
example, our own Moon is less than 1/80 the mass of Earth.

Of course, even a small moon can appear very large in a world’s sky, if it orbits quite closely. Our Moon
may originally have formed just outside Earth’s Roche limit, at about 2-3 times Earth’s radius. However,
as soon as a moon forms, tidal interactions will begin to push it outward. In our own case, the Theia
impact seems to have imparted a great deal of angular momentum (“spin”) to Earth; the rotation period
immediately after impact seems to have been as little as five hours. Tidal interactions quickly
transferred some of this angular momentum to the Moon, causing its orbit to rise to several times its
original radius within the first hundred million years. Today, of course, the Moon’s orbit is at about 60
times Earth’s radius. Any other terrestrial planet acquiring a major satellite due to a large impact is also
likely to see that satellite retreat to a significant distance.

It’s for these reasons that the standard design sequence won’t generate large, close-in moons based on
the large-impact process – the situation doesn’t appear to be stable on long timescales.

It might be reasonable to place an unusually large moon closer than the usual lower limit of 40 planetary
radii if the system is very young, or if the parent planet becomes tide-locked to its satellite early in the
process.
In fact, there is an example in our own planetary system. The dwarf planet Pluto and its major satellite
Charon appear to have formed due to a large impact (although the mechanism appears different than
that of the Earth-Moon pair). Pluto and Charon are tide-locked to each other. The ratio of their masses is
only about 8:1, and Charon’s orbital radius is only about 16 times that of Pluto. Charon appears about
3.8° wide in the Plutonian sky, several times the apparent size of the Moon from Earth.

There is another mechanism which might give rise to a very large “satellite” in the sky. This is the
appearance of a double planet or binary planet. In this case, rather than having a single planet coalesce
in a specific position, a pair of planetary-mass objects form, sharing the same orbit around the primary
star while also circling around their shared center of mass. If the two objects become tide-locked to one
another during the process of formation, they will not be subject to the tidal forces that tend to push a
single planet and its large-impact satellite apart.

Double planets of this type are (so far) an entirely theoretical notion; we’ve never detected an example,
and the usual models for planetary formation suggest such an event is unlikely. Still, double planets
often appear in science fiction, so you may wish to place one as a rare case.

To do this within the standard design sequence, begin in Step Eleven with a single planet forming in the
inner planetary system, whose formation orbit has not been disrupted by the migration of any dominant
gas giant.

Take the mass of the original planet and divide it into two portions, not necessarily evenly. To do this at
random, roll 3d6, divide by 20, and multiply by the original planet’s mass. The result is the mass of one
component of the binary planet; the other component will get the remainder. Remember that the
minimum mass for an inner-system planet (or one of a binary planet’s components) is 0.03 Earth-
masses.

Determine the eccentricity of the binary planet’s orbit around the primary star normally in Step Twelve.
Determine each component’s density, radius, and surface gravity normally in Step Thirteen.

Rather than test for natural satellites in Step Fourteen, treat the less massive component as an
unusually large natural satellite for the larger; neither component will have any natural satellites or
moonlets of its own. To determine the mutual orbital radius at random, roll 2d6+3 and multiply by the
radius of the larger component. The eccentricity of the mutual orbit will be very small, no more than
about 0.05.

In Step Sixteen, the two components will automatically be in a 1:1 spin-orbit resonance with one
another (that is, they will be tide-locked). In Step Seventeen, the two components will have the same
obliquity with respect to their orbital plane around the primary star; roll 3d6-8 (minimum 0) and take
the result as the obliquity in degrees.

For the rest of the design sequence, each component of the binary planet can be treated as an
independent world.

Example: Alice is designing a binary planet for one of her planetary systems. The original planet had a
planetesimal mass of 2.0 Earth-masses in Step Eleven, so she divides this exactly in half and ends up with
two components of almost exactly Earth’s mass. Deciding that each component has the same density
and radius as Earth, she generates their mutual orbital radius as about 51,000 kilometers. Each
component appears about 8.2° wide in the other’s sky.

Habitable Moons for Gas Giants


Many science fiction settings involve Earthlike worlds which appear as major satellites of gas giant
planets. This seems unlikely under our current models of planetary formation, which rarely produce
moons that are more than a small fraction of Earth’s mass. This fits the experience of our own planetary
system, where the major satellites of our own gas giants seem to mass no more than about one ten-
thousandth as much as their primary planet.

On the other hand, it's possible for a gas giant planet to capture a moon after the formation era. For
example, it seems likely that in our own planetary system, Neptune captured its sole major satellite
(Triton). It would be reasonable to give the occasional gas giant a captured moon, much larger than the
usual mass for its natural satellites. If you would like to create such a situation as a rare case, check the
following guidelines:

• The main planet in the relationship should be the system’s dominant gas giant.
• The dominant gas giant must have migrated inward from its original formation orbit, eliminating
at least one inward formation orbit in the process. One of the eliminated formation orbits must
have possessed enough planetesimal mass to produce a planet of the desired size.
• No Grand Tack should have taken place.
• The dominant gas giant must have the possibility of major satellites forming by natural accretion
(that is, its Hill radius must be at least 300 times its own radius).

In this case, the unusually massive moon will have the mass of one of the planets that would otherwise
have formed in an eliminated formation orbit.

Give the captured planet an orbital radius between 10 and 25 times that of the gas giant (roll 3d6+7 to
generate this at random). Roll 1d6: on a 4-6 the captured planet’s orbit will be retrograde (that is, it will
orbit in the opposite direction from most bodies in the planetary system). In this case, the captured
planet will have eccentricity of 0. Otherwise, the eccentricity will still be very low (0.05 or less). Captured
planets will be tide-locked to the primary gas giant.

The arrival of the captured planet is very likely to disrupt the orbits of any of the gas giant’s natural
satellites, most likely causing them to fall into the primary or ejecting them from the system entirely.
Determine how many natural satellites the gas giant would have had normally in Step Fourteen, then
roll 1d6 for each; only on a result of 6 will the natural satellite survive. Surviving natural satellites will
most likely be those closest to the gas giant.

World History
TO BE COMPLETED

You might also like