This is one of the webpages of Libarid A. Maljian at the Department of Physics at CSLA at NJIT.
New Jersey Institute of Technology
College of Science and Liberal Arts
Department of Physics
Introductory Astronomy and Cosmology, Section 101
Fall 2024
Fourth (Final) Examination lecture notes
Our Galaxy, the Milky Way Galaxy
A galaxy is a collection of
billions of star systems held together by gravity. Our Milky Way Galaxy is composed of roughly
one hundred billion star systems, and our Solar System is just one of these one
hundred billion star systems that together compose our Milky Way Galaxy. Since most star systems are binary star
systems, our Milky Way Galaxy is composed of roughly two hundred billion
stars. If we regard the average mass of
a star to be roughly 1M☉ (one
solar mass), we might suspect that the total mass of our Milky Way Galaxy is
roughly one or two hundred billion solar masses. However, the total mass of our Milky Way
Galaxy is much larger, roughly one trillion solar masses, for reasons we will
discuss shortly. Most of the star
systems that compose our Milky Way Galaxy are arranged
into the shape of a flat disk with spiral arms, making the shape of the Milky
Way Galaxy rather like a pinwheel.
Therefore, the Milky Way Galaxy is classified
as a spiral-disk galaxy. The galactic
disk is roughly thirty kiloparsecs (roughly one
hundred thousand light-years) in diameter; therefore, the galactic disk is
roughly fifteen kiloparsecs (roughly fifty thousand
light-years) in radius. Our Solar System
is in the galactic disk roughly halfway out from the center of our Milky Way
Galaxy, making our Solar System roughly eight kiloparsecs
(roughly twenty-five thousand light-years) from the center of our Milky Way
Galaxy. The galactic disk is only
roughly three hundred parsecs (roughly one thousand light-years) thick. Therefore, the galactic disk has a diameter
roughly one hundred times its own thickness; that is, the thickness of the
galactic disk is roughly one-hundredth of its own diameter. More plainly, the galactic disk is very
thin. Our Solar System is roughly midway
within the thickness of the galactic disk, being roughly one hundred and fifty
parsecs (roughly five hundred light-years) from either the top edge or the
bottom edge of the galactic disk. The
galactic bulge is a collection of stars arranged into
a more rounded shape near the center of our Milky Way Galaxy. If we could observe our Milky Way Galaxy edge
on, its galactic disk together with its galactic bulge would give it the
appearance of a fried egg. There are
some stars around the galactic disk (above and below the galactic disk); this is called the galactic halo.
The stars of the Milky Way
Galaxy move along giant orbits around the center of the Milky Way Galaxy. The star systems within the galactic disk,
including our own Solar System, orbit the center of our Milky Way Galaxy in
roughly the same angular (orbital) direction.
Therefore, we may regard our Milky Way Galaxy as rotating due to the
orbits of all of its individual star systems, but this rotation is certainly
not rigid body rotation. A galaxy is not
a solid object; a galaxy is composed of billions of individual star systems
each on their own orbit around the center of the galaxy. Therefore, the shape of a galaxy continuously
changes as all of its star systems move along their
individual orbits. Of course, we would
need to wait millions of years to notice significant changes in the shape of a
galaxy. Nevertheless, over timescales of
millions of years, we would observe that the spiral arms of a spiral-disk
galaxy are not permanent structures. The
stars closer to the center of a galaxy take less time to complete one orbit,
since they have less distance to travel.
The stars further from the center of a galaxy take more time to complete
one orbit, since they have more distance to travel. Therefore, if we could observe a spiral-disk
galaxy over timescales of millions of years, we would see its shape
continuously change as some spiral arms disperse and new spiral arms coalesce
depending on where all of the star systems happen to be located on their
individual orbits. There are also
variations in the concentration of stars throughout the Milky Way Galaxy. Stars are somewhat closer together in some
regions of our Milky Way Galaxy, while stars are somewhat further apart in
other regions of our Milky Way Galaxy.
Moreover, these variations in stellar concentrations propagate
throughout all the stars that together form our Milky Way Galaxy. These propagating disturbances are called spiral density waves, since they contribute to
the changing structure of the spiral arms of the Milky Way Galaxy. These spiral density waves also cause
variations in the density of gases within the Milky Way Galaxy, as we will
discuss shortly. Since galaxies do not
suffer from rigid body rotation, it is not meaningful to ask how long it takes
a galaxy to complete one rotation, since each star system is on its own orbit,
each taking a different amount of time to orbit around the center of the
galaxy. We could however ask for the
average duration of time it takes a galaxy to complete one rotation. Since our own Solar System is roughly halfway
out from the center of our Milky Way Galaxy, we may regard the time it takes
our Solar System to complete one orbit around the center of our Milky Way
Galaxy as the average duration of time it takes our Milky Way Galaxy to
complete one rotation. This is called one galactic year, and it is between two hundred
million years and two hundred and fifty million years. Our Milky Way Galaxy is roughly ten billion
years old, as we will discuss shortly.
Therefore, our Milky Way Galaxy has not completed a large number of
rotations over its entire history. Ten
billion divided by two hundred and fifty million is only forty, and ten billion
divided by two hundred million is only fifty.
Therefore, our Milky Way Galaxy has only rotated between forty and fifty
times over its entire ten-billion-year history.
Our Solar System is only roughly five billion years old, roughly
one-half of the age of the Milky Way Galaxy.
Therefore, our Solar System has only completed between twenty to
twenty-five orbits around the center of the Milky Way Galaxy over its own
particular history.
The study of the structure
and the evolution of galaxies is called galactic
dynamics, and astrophysicists who study the structure and the evolution of
galaxies are called galactic dynamicists. Both the structure and the evolution of
galaxies is studied through the orbital motions of the
individual star systems that compose a galaxy.
Theoretical galactic dynamicists program a
computer with billions of stars attracting each other gravitationally. While running the simulation, the computer
can display various images of the positions of the stars over the course of the
simulated time. In this way, theoretical
galactic dynamicists can study the changing structure
(the changing shape) of a galaxy.
Observational galactic dynamicists collect
light from other galaxies. Since most of
the star systems of a spiral-disk galaxy rotate in roughly the same angular
(orbital) direction, the stars on one side of a spiral-disk galaxy will happen
to be moving toward us while the stars on the other side of the spiral-disk
galaxy will happen to be moving away from us.
Thus, the light from one side of the galaxy will be blueshifted
relative to its galactic center, and the light from the other side of the
galaxy will be redshifted relative to its galactic center. By measuring these blueshifts
and redshifts, we can calculate the speed with which the stars are moving
within the galaxy, and galactic dynamicists can then
predict the structure (the shape) the galaxy will have in the future as well as
the structure (the shape) the galaxy had in the past.
By observing the orbits of
stars near the center of our Milky Way Galaxy and by calculating the
gravitational force acting on these stars, galactic dynamicists
have determined that there is a supermassive black hole at the center of our
Milky Way Galaxy. A supermassive black
hole has a mass at least in the millions of solar masses. Presumably, a supermassive black hole was
initially born a stellar black hole from the Type II supernova of a very high
mass star. Over billions of years, the
stellar black hole continuously devoured all of the gas around it and thus
continued to grow in mass, eventually becoming a supermassive black hole. By observing the orbits of stars near the
center of other galaxies and by calculating the gravitational force acting on
them, galactic dynamicists have determined that there
is a supermassive black hole at the center of every major galaxy in the
universe. Much more recently,
astronomers have actually succeeded in imaging these supermassive black holes. Using many radio telescopes working together
as a single interferometer, astronomers have produced radio images of the black
event horizon of supermassive black holes against the gases of the surrounding
space. In the year 2019, astronomers
imaged the event horizon of the supermassive black hole at the center of a
galaxy roughly sixteen megaparsecs (roughly fifty
million light-years) distant. In the
year 2022, astronomers imaged the event horizon of the supermassive black hole
at the center of our own Milky Way Galaxy.
Although supermassive black holes are incredibly massive as compared
with stellar black holes, they still have a small mass as compared with the
mass of an entire galaxy. Although the
gravity of a supermassive black hole is strong enough to determine the orbits
of stars near a galactic center, the gravity of a supermassive black hole is
nowhere nearly strong enough to hold an entire galaxy together. For example, the supermassive black hole at
the center of our Milky Way Galaxy is roughly four million solar masses, but
the mass of our entire Milky Way Galaxy is roughly one trillion solar
masses. Thus, the mass of the
supermassive black hole at the center of our Milky Way Galaxy is roughly four
ten-thousandths of one percent of the mass of the entire Milky Way Galaxy. This tiny fraction (tiny percentage) should
be contrasted with our Solar System, where the mass of our Sun is roughly 99.9
percent of the mass of the entire Solar System.
Thus, the Sun’s gravity overwhelmingly dominates our Solar System, and so it is indeed the Sun’s gravity that holds our Solar System
together. Since the mass of a
supermassive black hole is such a tiny fraction (tiny percentage) of the mass
of an entire galaxy, supermassive black holes do not hold galaxies
together. Something much more mysterious
than black holes is responsible for holding galaxies together.
A rotation curve is a graph
of the orbital speed of stars around the center of a galaxy as a function of
their distance from the center of the galaxy.
By measuring the orbits of stars far from the center of our Milky Way
Galaxy, galactic dynamicists have determined that our
Milky Way Galaxy’s rotation curve flattens far from the galactic center. In fact, the rotation curve of all
spiral-disk galaxies flattens far from their galactic center. It is not difficult to use the mathematical
laws of gravitation to determine the distribution of mass that would exert the
gravitational force necessary to cause a flat rotation curve. We conclude that the billions of star systems
that compose a galaxy only account for roughly one-tenth (roughly ten percent)
of the mass of an entire galaxy. Roughly
ninety percent of the mass of a galaxy is distributed
throughout an enormous sphere surrounding all the star systems of the
galaxy. As we discussed earlier in the
course, we determine the composition of anything in the universe such as
planets, stars, and nebulae through spectroscopy, the measuring of the spectrum
of the light we receive from the object.
Whether the light has an absorption spectrum or an emission spectrum, we
consult tabulated spectral lines to determine which atoms absorbed or emitted
those wavelengths of light. However, we
receive no light whatsoever from anywhere across the entire Electromagnetic Spectrum
from the enormous distribution of mass dominating galaxies. All atoms interact with photons, and atoms
are composed of protons, neutrons, and electrons. Protons and electrons certainly interact with
photons, since they have electric charge.
Although neutrons are neutral, they nevertheless also interact with
photons, since neutrons still have electromagnetic properties. If the enormous distribution of mass
dominating galaxies does not interact with any photons at all, then this
dominating mass cannot be composed of atoms, nor can this dominating mass be
composed of the constituents of atoms (protons, neutrons, or electrons). Therefore, astronomers have no idea what
composes this mysterious mass that dominates galaxies. In their complete ignorance, astronomers use
the term dark matter for this mysterious mass.
Therefore, roughly ninety percent of the mass of our Milky Way Galaxy is
dark matter distributed over an enormous sphere surrounding all the star
systems of our Milky Way Galaxy. This
enormous sphere is called the dark matter halo. Since every spiral-disk galaxy in the
universe has a flat rotation curve, we deduce that roughly ninety percent of
the mass of every spiral-disk galaxy in the universe is
distributed over a mysterious dark matter halo surrounding the star
systems of the galaxy. As we will
discuss shortly, dark matter in fact composes roughly ninety percent of the
mass of all galaxies in the universe, not just spiral-disk galaxies. Therefore, dark matter composes roughly
ninety percent of the mass of the entire universe. The remaining roughly ten percent of the mass
of the universe that is composed of atoms is called
normal matter. However, if dark matter
composes roughly ninety percent of the mass of the universe, perhaps dark
matter should be renamed as normal matter; it is stars and
planets and mountains and buildings and humans and mobile telephones
that are composed of abnormal (atomic) matter!
As we will discuss shortly, there is further evidence in addition to
flat rotation curves for the predominance of dark matter in the universe. The ten percent normal (atomic) mass of the
universe is composed overwhelmingly of stars and gas. Planets, moons, asteroids, and comets compose
a tiny fraction (tiny percentage) of this normal (atomic) mass. Therefore, we will often refer to normal
(atomic) mass as luminous mass, since stars are luminous. We now summarize the composition of matter in
the universe. Roughly ninety percent of
the mass of the universe is composed of dark matter. The remaining roughly ten percent of the mass
of the universe is composed of normal (atomic) matter. Roughly three-quarters (roughly seventy-five
percent) of this roughly ten percent normal (atomic) matter is hydrogen,
roughly one-quarter (roughly twenty-five percent) of this roughly ten percent
normal (atomic) matter is helium, and all the other atoms on the Periodic Table
of Elements compose a tiny fraction (tiny percentage) of this roughly ten
percent normal (atomic) matter.
It is unsettling to discover
that we do not know the composition of roughly ninety percent of the mass of
the universe. We are certain that dark
matter has mass that exerts gravitational forces in the same way as normal
(atomic) matter; this is how we discovered dark matter in the first place. Other than mass that exerts normal
gravitational forces, what is dark matter exactly? One theory claims that dark matter is
actually composed of normal matter.
Brown dwarf stars are very cool and dim, as we discussed earlier in the
course. According to this theory, the
dark matter is actually an enormous number of brown dwarf stars in the galactic
halo surrounding all the other stars of a galaxy. Consequently, these brown dwarf stars are called massive astrophysical compact halo objects, which
astrophysicists always abbreviate as MACHOs. This MACHO theory is probably not
correct. Although brown dwarf stars
radiate very little visible light, they radiate a fair amount of infrared light
at their cool temperatures. To account
for the predominance of dark matter, we would need such an enormous number of
brown dwarf stars distributed throughout the galactic halo that all of their
infrared light would add to a significant amount. However, we do not receive any infrared light
or any other type of photons whatsoever from dark matter. Moreover, cosmological observations and
cosmological calculations place constraints upon the amount of normal matter
that formed shortly after the Big Bang that created the universe, as we will
discuss shortly. To account for the
predominance of dark matter, the number of brown dwarf stars would exceed these
cosmological constraints upon the amount of normal matter that fills the
universe. Therefore, many
astrophysicists agree that dark matter cannot be composed of normal (atomic)
matter. An opposing theory to explain
the composition of dark matter claims that dark matter is composed of exotic
quantum-mechanical particles. There is a
highly speculative theory called Supersymmetric Relativistic Quantum Field Theory,
or Supersymmetry for short. According to
Supersymmetric Relativistic Quantum Field Theory, for every particle of matter
or antimatter in the universe, there is a corresponding supersymmetric
particle. For example, electrons,
positrons (antielectrons), neutrinos, and antineutrinos are
all classified as leptons, but Supersymmetric Relativistic Quantum Field
Theory claims that there are supersymmetric leptons called sleptons. This speculative theory also claims that
there are supersymmetric quarks called squarks,
supersymmetric photons called photinos,
supersymmetric gluons called gluinos, and
supersymmetric gravitons called gravitinos. Other supersymmetric particles include winos
and zinos. No
supersymmetric particle has ever actually been observed. In other words, sleptons,
squarks, photinos, gluinos, gravitinos, winos, and zinos are all purely hypothetical particles. However, subatomic particle accelerators may
be able to create these supersymmetric particles, as we will discuss
shortly. Although Supersymmetric
Relativistic Quantum Field Theory is highly speculative, some astrophysicists
believe that dark matter is composed of these supersymmetric particles. These supersymmetric particles must have a
significant amount of mass to account for the predominance of dark matter in
the universe. Therefore, these
hypothetical supersymmetric particles are also called
weakly interacting massive particles, which astrophysicists always abbreviate
as WIMPs. To
summarize these two opposing theories to explain the composition of dark
matter, some astrophysicists believe that the dark matter is composed of MACHOs (massive astrophysical compact halo objects) while
other astrophysicists believe that the dark matter is composed of WIMPs (weakly interacting massive particles), and more
astrophysicists side with the WIMPs over the MACHOs!
It is a common misconception
that outer space is perfect vacuum, but there is no such thing as perfect
vacuum. In fact, a perfect vacuum would
violate the laws of physics. Outer space
is actually filled with very diffuse gas called the
interstellar medium, which astrophysicists always abbreviate ISM. The interstellar medium is concentrated
within the galactic disk, filling the space between stars, hence its name. The interstellar medium is composed of
roughly three-quarters (roughly seventy-five percent) hydrogen, roughly
one-quarter (roughly twenty-five percent) helium, and tiny amounts of all the
other atoms on the Periodic Table of Elements.
We determine the composition of the interstellar medium through spectroscopy. We find absorption lines in the starlight
that passes through the interstellar medium.
By measuring the wavelengths of these absorption lines and consulting
tables of spectra, we can determine which atoms absorbed these spectral lines
and thus determine the composition of the interstellar medium. We also find emission lines within the light
from the interstellar medium. Again, by
measuring the wavelengths of these emission lines and consulting tables of
spectra, we can determine which atoms emitted these spectral lines and again
determine the composition of the interstellar medium. The gases within the
interstellar medium are pushed by many different forces, including thermal
pressures, gravitational forces, magnetic pressures, and even cosmic rays (ultra high-energy particles). All these different forces are comparable in
strength with each other in interstellar space (the space between star
systems). Thus, the gases within the
interstellar medium are pushed in seemingly random directions, causing some
regions within the interstellar medium to be more dense than average and other
regions within the interstellar medium to be less dense (or more tenuous) than
average. A region of the interstellar
medium that is more dense than average is called a diffuse nebula, since even
these more dense regions of the interstellar medium are still diffuse (low
density) by human standards. The gases
within these diffuse nebulae are sufficiently cool that they radiate more infrared
light and less visible light. Consequently,
infrared images of a diffuse nebula typically reveal its gases much more
clearly than optical images. Within a
diffuse nebula, gases are pushed by many different forces
that are all comparable in strength with each other. Hence, the gases within a diffuse nebula are pushed in seemingly random directions, causing
variations in density even within a diffuse nebula. Small regions within a diffuse nebula may
become dense enough that gravity dominates over all the other forces. Thus, those small regions of the diffuse
nebula will collapse from their self-gravity (under their own weight),
eventually becoming star systems. Note
however that the stars born within a diffuse nebula provide heat and thus
thermal pressures that may balance or even exceed gravitational forces. Moreover, the radiation pressure from the
light radiated from the stars born within a nebula will push the surrounding
gases outward. The stellar winds from
the stars born within a nebula will also push the surrounding gases outward. As a result of
thermal pressures, radiation pressures, and stellar winds from the stars that
are born within a diffuse nebula, most of the gases of a diffuse nebula will
not form stars; only a tiny fraction of the total mass of a diffuse nebula will
form stars. On the other hand, the
energy from a nearby supernova may compress the gases of a particular region of
the interstellar medium to sufficient densities for gravity to dominate over
all other forces, thus inducing star formation.
A spiral density wave may also compress the gases of a particular region
of the interstellar medium to sufficient densities for gravity to dominate over
all other forces, thus inducing star formation.
Since the interstellar medium is concentrated within the galactic disk,
this is where star formation occurs in the Milky Way Galaxy. Stars are continuously born within the spiral
arms of the galactic disk of the Milky Way Galaxy. There is virtually no star formation in the
galactic halo, the region around (above and below) the galactic disk, since
there are virtually no interstellar gases outside of the galactic disk. Therefore, the stars in the galactic halo
must be relatively old, since without active star formation there would be no
newly born stars. The stars in the
galactic disk must be relatively young, since stars are continuously born from
diffuse nebulae within the interstellar medium within the galactic disk. Astronomers have named the stars within the
galactic disk Population I (Roman numeral) stars, and astronomers have named
the stars within the galactic halo Population II (Roman numeral) stars. Again, Population I stars are comparatively
young, while Population II stars are comparatively old. Since star formation is still active in the
galactic disk, some newly born stars will be high mass, hot, luminous, and blue
(early-type) main sequence stars. Such
stars have short lifetimes ending with violent supernovae, as we discussed
earlier in the course. A supernova
explosion synthesizes all the atoms across the entire Periodic Table of
Elements and throws them into the surrounding interstellar medium through the
hot and rapidly expanding supernova remnant, as we discussed earlier in the
course. Thus, the interstellar medium is polluted or enriched with these new nuclei, causing future
diffuse nebulae to be similarly polluted or enriched. Hence, Population I stars have comparatively
higher mass fractions of these nuclei.
Since Population II stars are comparatively old, they must be low mass,
cool, dim, and red (late-type) main sequence stars, since these stars have
longer lifetimes. These old stars were
born when the universe was younger; therefore, there was less time for high
mass stars to synthesize heavier elements.
Thus, Population II stars have comparatively lower mass fractions of
these heavier nuclei. Recall that the
normal (atomic) mass of the universe is roughly three-quarters (roughly
seventy-five percent) hydrogen, roughly one-quarter (roughly twenty-five percent)
helium, and tiny amounts of all the other atoms on the Periodic Table of
Elements. Therefore, astrophysicists
place all the atoms on the Periodic Table of Elements into only three
categories: hydrogen, helium, and metals.
In other words, astrophysicists use the word metal for any atom besides
hydrogen or helium. Many students are
offended by this categorization, claiming that oxygen and
nitrogen and neon for example are not metals. In a chemistry course, this is certainly the
case. Nevertheless, astrophysicists
classify all atoms besides hydrogen and helium as metals. The metallicity of a star or a nebula or even
an entire galaxy is the fraction (percentage) of its normal (atomic) mass that is composed of metals (all atoms besides hydrogen and
helium). Population I stars have relatively
higher metallicities, since they are a later (more recent) generation of stars
polluted or enriched by metals from the supernova explosions of an earlier
generation of high-mass stars. Caution:
by higher metallicity we mean at most one percent or two percent; all stars in
the universe are composed of roughly three-quarters (roughly seventy-five
percent) hydrogen, roughly one-quarter (roughly twenty-five percent) helium,
and only tiny amounts of metals. Population
II stars have relatively lower metallicities, since they are an older
generation of stars that were not significantly polluted or enriched by metals,
since they were born when the universe was younger and therefore there was less
time for high mass stars to synthesize metals.
The metallicities of Population II stars is roughly one-tenth of one
percent, roughly ten times smaller than the metallicities of Population I
stars. That is, the metallicities of
Population I stars is roughly ten times greater than the metallicities of
Population II stars. The very first
generation of stars born in the entire universe had zero metallicity, since
they were composed of pure hydrogen and helium since there was no earlier
generation of stars to synthesize any metals.
These stars are called Population III (Roman
numeral) stars. In other words,
Population III stars should really be renamed first-generation stars,
Population II stars should really be renamed second-generation stars, and
Population I stars should really be renamed third-generation stars! Nevertheless, we will continue to use the
standard Roman numeral designations. In
summary, Population I stars are within the galactic disk and are comparatively
high mass, hot, luminous, blue, young, high-metallicity stars, while Population
II stars are within the galactic halo and are comparatively low mass, cool,
dim, red, old, low-metallicity stars.
Since our Sun is in the galactic disk, we might suspect that our Sun is
a Population I star. Indeed, the
metallicity of our Sun is between one percent and two percent. Moreover, the very existence of four
terrestrial planets that formed with the Sun composed of metals such as iron,
nickel, silicon, oxygen, and nitrogen is direct evidence that our Sun is indeed
a Population I star.
As we discussed earlier in
the course, the main sequence is a population-abundance sequence. That is, most stars are born late-type main
sequence stars (cool, dim, red, and low-mass with long
lifetimes), while few stars are born early-type main sequence stars (hot,
luminous, blue, and high-mass with short lifetimes). If the very first generation of stars born in
the entire universe, Population III stars with zero metallicity, formed in the
same way that stars continue to form today, then most of them would still
remain to the present day, since most stars are born low-mass with long
lifetimes. However, no Population III
stars with zero metallicity have ever been discovered. This suggests that the first generation of
stars born in the universe formed by a mechanism different from later star
formation mechanisms. Although
astrophysicists continue to debate the mechanism by which Population III stars
formed, it seems that we must conclude that all Population III stars were born
early-type main sequence stars (hot, luminous, blue, and high-mass
with short lifetimes). This would
explain why there are no Population III stars remaining in the universe today;
all of them were born high-mass main sequence stars with short lifetimes and
hence all of them died within only millions of years after their birth. As we will discuss shortly, there is further
evidence that all Population III stars were born high-mass stars with short
lifetimes.
All stars are born in
clusters, since many stars are born within a diffuse nebula
simultaneously. However, most stars do
not remain in clusters indefinitely.
After a star cluster is born from a diffuse nebula, the individual stars
drift apart from one another as they move along their own orbital trajectories
through our Milky Way Galaxy. Therefore,
most stars are not members of star clusters.
For example, our Sun is not presently a member of a star cluster,
although our Sun was presumably born a member of an ancient star cluster that
has long since dispersed. Star clusters
within the galactic disk that are composed of Population I stars are called
open star clusters. The Pleiades Cluster
in the constellation Taurus (the bull) and the Ptolemy Cluster in the
constellation Scorpius (the scorpion) are beautiful examples of open star
clusters. Star clusters within the
galactic halo (outside of the galactic disk) that are composed of Population II
stars are called globular star clusters. The Hercules Cluster in the constellation
Hercules (the hero) and Omega Centauri in the constellation Centaurus (the
centaur) are beautiful examples of globular star clusters. The closest star cluster to our Solar System
is the Hyades Cluster, an open star cluster within the galactic disk roughly
fifty parsecs (roughly 150 light-years) distant in the direction of the constellation
Taurus (the bull). However, there are
several other groups of stars even closer to us than the Hyades Cluster. These stellar groups were probably born as
open star clusters, and they are currently in the process of dispersing. The closest such stellar group is the Ursa Major Stellar Group located within the galactic disk
roughly twenty-five parsecs (roughly eighty light-years) distant in the
direction of the constellation Ursa Major (the big
bear).
Everything we have discussed
about Population I stars and Population II stars applies to open star clusters
and globular star clusters, respectively.
In particular, open star clusters are composed of comparatively high
mass, hot, luminous, blue, young, high-metallicity stars as compared with
globular star clusters which are composed of comparatively low mass, cool, dim,
red, old, low-metallicity stars. As we
discussed earlier in the course, the Hertzsprung-Russell
diagram of a star cluster reveals the history of the cluster. For example, we can calculate the age of a
star cluster from the main-sequence turnoff on its Hertzsprung-Russell
diagram. We know that open star clusters
are young, since the main-sequence turnoff on their Hertzsprung-Russell
diagrams is early. Conversely, we know
that globular star clusters are old, since the main-sequence turnoff on their Hertzsprung-Russell diagrams is late. In fact, globular star clusters are the
oldest organizations in the Milky Way Galaxy; many globular star clusters are
roughly ten billion years old. This is
how we know the age of the entire Milky Way Galaxy, from the age of its oldest
organizations. Globular star clusters
contain an abundance of white dwarfs, since globular star clusters are old and
hence there has been sufficient time for many low-mass stars in the cluster to
live their long lifetimes and reach the very end of their evolutions, finally
ending their lives as white dwarfs.
Conversely, open star clusters contain few white dwarfs, since open star
clusters are young and hence there has only been sufficient time for very few
of the low-mass stars in the cluster to live their lifetimes and reach the very
end of their evolutions, finally ending their lives as white dwarfs. If an open star cluster is particularly
young, there may be no white dwarfs within the cluster at all, since there has
only been sufficient time for high-mass stars in the cluster to live their
short lifetimes and reach the very end of their evolutions, finally ending
their lives with supernova explosions and leaving behind neutron stars or black
holes. Particularly young open star
clusters are often still embedded within the diffuse
nebula from which they formed. These are called embedded star clusters.
Open star clusters are irregularly shaped, hence the term open. Globular star clusters are
spherically shaped, hence the term globular. Open star clusters are within the galactic
disk orbiting the center of the Milky Way Galaxy together with most of the
stars that compose the Milky Way Galaxy.
Therefore, many open star clusters move at slow speeds relative to our
Solar System, since we are actually moving together in roughly the same angular
(orbital) direction at roughly the same speed.
Globular star clusters move at fast speeds relative to our Solar System,
since they are within the galactic halo (outside of the galactic disk) moving
along random orbits around the center of the Milky Way Galaxy. Open star clusters typically contain only
several hundred stars. Consequently, the
mutual gravitational attraction of all the stars within an open star cluster is
insufficient to hold the cluster together.
Thus, stars are not gravitationally bound to each other within an open
cluster, and therefore the stars within an open star cluster will disperse from
one another within only several million years.
Globular star clusters typically contain hundreds of thousands of
stars. Consequently, the mutual
gravitational attraction of all the stars within a globular star cluster is
sufficient to hold the cluster together.
Thus, stars are gravitationally bound to each other within a globular
star cluster, and therefore the stars within a globular star cluster will not
disperse from one another.
Just as diffuse nebulae are
regions of the interstellar medium that are more dense than average, bubbles
are regions of the interstellar medium that are less dense than average. As we discussed earlier in the course, the
hottest and most luminous main sequence stars have spectral type either O or B. An O-type or a B-type star is so luminous
that the strong radiation pressure from its light will push gases within the
interstellar medium away from the star.
The stellar wind of an O-type or B-type star is also sufficiently strong
to push gases within the interstellar medium away from the star. The result is a bubble: a spherical region
around the O-type or B-type star where the interstellar medium is less dense
than average. Every O-type or B-type
star has a bubble surrounding it. Star
clusters with a significant number of O-type and B-type stars are called OB associations.
The combined luminosities and the combined stellar winds from all the
O-type and B-type stars within these OB associations push the gases of the
interstellar medium so strongly that the entire OB
association is surrounded by a superbubble. Every OB association has a superbubble surrounding it.
These superbubbles are enormous, hundreds of
light-years across. The galactic disk is
only roughly one thousand light-years thick, as we discussed. Therefore, a superbubble
can grow to sufficient size to burst out of the galactic disk, ejecting
material out of the galactic disk. The
gravity of the galactic disk does pull this material back toward the galactic
disk however, and the subsequent collision of this ejected material with the gases
of the interstellar medium within the galactic disk may induce star
formation. As we discussed earlier in
the course, O-type and B-type stars live short main-sequence lifetimes and die
with a violent supernova. It is not
difficult to calculate that the supernova remnant ejected by a supernova at
first moves so fast that its gases should be able to escape from the
gravitational attraction of the entire Milky Way Galaxy. However, this rapidly expanding supernova
remnant soon collides with the gases of the surrounding interstellar medium,
and the collision slows the expanding supernova remnant to speeds slower than
the galactic escape speed, thus keeping the gases within the galactic
disk. Moreover, the collision compresses
the surrounding gases of the interstellar medium. These gases may be
compressed to sufficient densities for gravity to dominate over other
forces within small parts of the surrounding interstellar medium, thus inducing
star formation. In summary, high mass
stars trigger the birth of new stars, through superbubbles
that burst out from and then fall back toward the galactic disk and through the
collision of supernova remnants with the surrounding interstellar medium.
There are three different
types of diffuse nebulae: absorption nebulae, emission nebulae, and reflection
nebulae. Absorption nebulae tend to
appear black in color, since they absorb photons. Note however that an infrared image of an
absorption nebula often clearly reveals the gases within it. The Horsehead Nebula in the constellation
Orion (the hunter) is a beautiful example of an absorption nebula. Other diffuse nebulae have stars within them
that were recently born from the gases within the nebula itself. The light radiated by these
stars within the nebula are absorbed by the surrounding gases of the
nebula. This transitions the electrons
within the atoms composing the nebula to higher-energy quantum states. The electrons then transition back down to
lower-energy quantum states, thus emitting photons. The result is an emission nebula. All nebulae are composed of mostly hydrogen
gas, and there is a particular photon within the emission spectrum of the
hydrogen atom that falls within the red part of the visible light spectrum,
causing emission nebulae to often appear red in
color. Note however that an emission
nebula often displays a variety of colors, caused by photons emitted from other
transitions of electrons within the hydrogen atom and also
transitions of electrons within other atoms composing the nebula in addition to
hydrogen. There are also transitions
within the hydrogen atom and other atoms that result in the emission of
ultraviolet photons. Therefore, an
ultraviolet image of an emission nebula often more clearly reveals the gases
within it than a visible light image.
The Orion Nebula in the constellation Orion (the hunter) is a beautiful
example of an emission nebula.
Reflection nebulae tend to appear blue in color, since shorter
wavelengths of light are more preferentially scattered than longer wavelengths
of light, and blue is the short-wavelength end of the visible light
spectrum. The Witch Head Nebula in the
constellation Orion (the hunter) is a beautiful example of a reflection nebula. Note that most diffuse nebulae are a
combination of all three types (absorption, emission, and reflection), such as
the Trifid Nebula in the constellation Sagittarius
(the centaur archer).
The molecules of the Earth’s
atmosphere scatter our Sun’s light. The
shorter wavelengths are more preferentially scattered than the longer
wavelengths, causing the daytime sky to appear blue. The sky during sunrises and sunsets is red
for the same reason. Blue light has been
scattered out of sunlight, causing the daytime sky to appear blue. By the time sunlight has traveled through the
Earth’s atmosphere to arrive at someone on the cusp of the daytime side of the
Earth where it is sunrise or sunset, most of the short-wavelength blue light
has been subtracted from sunlight, leaving only the long-wavelength red light. The same effect occurs within the
interstellar medium. As light traverses
interstellar space, photons are absorbed and scattered by the
gases of the interstellar medium.
The shorter-wavelength blue light is more preferentially scattered,
leaving the longer-wavelength red light.
Astronomers call this effect reddening, although this term reddening is
misleading, since it implies that red light has been added
to the starlight. In actuality, blue
light has been subtracted; hence, astronomers should
rename this reddening effect as de-bluing instead! Nevertheless, all astronomers call this
effect reddening. As a
result of reddening, whenever we calculate the temperature of a star
using color indices as we discussed earlier in the course, we are actually
calculating an incorrect temperature for the star. The temperature we calculate is cooler than
the actual temperature of the star, since red light corresponds to cooler
temperatures. That is, the true
temperature of the star is hotter than our calculated temperature. Astronomers try to estimate the total amount
of reddening that has occurred while the starlight traversed interstellar space
through the interstellar medium.
Astronomers then add this total reddening (actually scattered blue
light) back into the color index calculations to determine the true temperature
of the star. More strictly, astronomers
must estimate the total extinction to determine the true temperature of the
star. Extinction is the total amount of
light that has been either scattered or absorbed by the interstellar
medium. The total extinction of light as
it traverses through the galactic disk is quite severe. After traveling just a few kiloparsecs (several thousand light-years) through the
interstellar medium of the galactic disk, one hundred percent extinction is attained, meaning that none of the starlight
remains! In other words, we cannot
observe visible light beyond a few kiloparsecs
(several thousand light-years) within our own galactic disk! Observing other galaxies beyond our Milky Way
Galaxy along the direction of our galactic disk is therefore hopeless. Fortunately, our galactic disk is rather
thin, as we discussed. Therefore,
extinction is less severe along directions perpendicular to the galactic
disk. Therefore, we are only able to
observe the extragalactic universe (the universe beyond our Milky Way Galaxy)
in directions perpendicular to our galactic disk, above and below the galactic
plane.
Although the galactic disk is
essentially opaque to visible light, astronomers use other wavelengths of light
to observe through our galactic disk, such as high-energy X-rays. Some low-energy photons are also able to
traverse through the galactic disk with minimal extinction. The primary photon that astronomers use to
observe through our galactic disk is twenty-one-centimeter-wavelength
photons. Hydrogen is the most abundant
atom composing the interstellar medium.
Hydrogen is also the simplest atom in the universe; it has only one
electron around its nucleus, and its nucleus is composed of a single
proton. As we discussed earlier in the
course, quantum-mechanical particles have an intrinsic angular momentum,
commonly known as spin. Within the
hydrogen atom, the spins of the proton and the electron can be either parallel
to each other or antiparallel to each other.
The antiparallel configuration is at a lower-energy quantum state than
the parallel configuration. If the spins
of the proton and the electron within the hydrogen atom happen to be parallel,
the spins may transition to the antiparallel configuration, thus decreasing the
energy of the entire atom. With this
transition from the parallel configuration to the antiparallel configuration,
the hydrogen atom emits a photon with a frequency of roughly 1420 megahertz and
a wavelength of roughly twenty-one centimeters.
Since roughly three-quarters (roughly seventy-five percent) of the
normal (atomic) mass of the universe is hydrogen atoms that continuously emit
twenty-one-centimeter-wavelength photons, the universe is filled with photons
with a wavelength of twenty-one centimeters.
These particular photons fall in the microwave band of the
Electromagnetic Spectrum, and microwaves suffer minimal extinction while
traversing through the interstellar medium.
Therefore, astronomers have determined the structure (the shape) of our
galactic disk by mapping the twenty-one-centimeter-wavelength photons emitted
by hydrogen atoms.
Again, our Milky Way Galaxy
is a spiral-disk galaxy, meaning that most of its star systems are arranged in a flat disk with spiral arms as they orbit
the center of our Milky Way Galaxy in roughly the same angular (orbital)
direction. The galactic disk has two
major spiral arms: the Perseus Arm and the Scutum-Centaurus-Crux Arm. There are a number of minor spiral arms, such
as the Norma Arm, the Carina-Sagittarius Arm, and the Orion Arm. As we discussed, our Solar System is roughly
halfway out from the center of our Milky Way Galaxy, roughly eight kiloparsecs (roughly twenty-five thousand light-years) from
the galactic center. More precisely, our
Solar System is within the minor Orion Arm, which is itself next to the major
Perseus Arm. From our location, the
galactic center is in the direction of the constellation Sagittarius (the
centaur archer). Astronomers have determined
the location of the galactic center from the distribution of globular star
clusters in the galactic halo. The
globular star clusters of our Milky Way Galaxy are roughly
spherically distributed throughout the galactic halo around the galactic
bulge. More precisely, the globular star
clusters are roughly spherically distributed around a
point roughly eight kiloparsecs (roughly twenty-five
thousand light-years) from our Solar System in the direction of the
constellation Sagittarius (the centaur archer).
The galactic disk is thin as compared with its diameter, as we
discussed. The flatness of our galactic
disk manifests itself in the sky. For
thousands of years, humans observed a band of milk around the entire sky they
called the milky way, as we discussed earlier in the course. Galileo Galilei used his primitive telescope
to discover that the milky way is not in fact milk; even a primitive telescope
reveals that the milky way is actually innumerable stars sufficiently crowded
together in the sky that with the naked eye all of their light blends together
so as to appear to be milk. Today we
realize that this milky way in the sky is actually our
flat galactic disk projected onto our sky.
When we observe into the direction of the milky way
in the sky, we are actually observing into our galactic disk. When we observe into the direction of the
milky way in the sky in the direction of the constellation Sagittarius (the
centaur archer), we are actually observing into our galactic disk and toward
our galactic center. When we observe
into the direction of the milky way in the sky in the opposite direction of the
constellation Sagittarius (the centaur archer), we are actually observing into
our galactic disk but away from our galactic center. This is often called the galactic anticenter, and it is in the direction of the milky way in the sky but in the opposite direction of the
constellation Sagittarius (the centaur archer).
More precisely, this galactic anticenter is
near the intersection of the three constellations Auriga (the charioteer),
Gemini (the twins), and Taurus (the bull).
If we observe directions off of the milky way
in the sky, we are actually observing along directions above or below our
galactic disk, which are the only directions where we may observe the
extragalactic universe, the universe beyond our Milky Way Galaxy.
Galactic Properties
There are a few dozen small
galaxies near our Milky Way Galaxy.
These are satellite galaxies, since they orbit our Milky Way
Galaxy. Caution: a satellite is anything
that orbits anything else. The Moon is a
satellite of the Earth, the Earth is a satellite of the Sun, and entire
galaxies can be satellites of other galaxies.
The two closest small satellite galaxies to our Milky Way Galaxy are the
Canis Major Dwarf Galaxy and the Sagittarius Dwarf
Galaxy. These galaxies are named for the constellations wherein they reside in our
sky, the constellation Canis Major (the big dog) and
the constellation Sagittarius (the centaur archer). The Canis Major Dwarf
Galaxy is roughly eight kiloparsecs (roughly
twenty-five thousand light-years) from our Solar System. The constellation Canis
Major (the big dog) lies close to the milky way in the
sky. Therefore, the Canis
Major Dwarf Galaxy is in physical contact with our galactic disk. The Sagittarius Dwarf Galaxy is roughly
twenty-five kiloparsecs (roughly seventy-five
thousand light-years) from our Solar System.
The constellation Sagittarius (the centaur archer) lies close to the milky way in the sky in the direction toward our galactic
center, as we discussed. Therefore, the
Sagittarius Dwarf Galaxy is on the other side of our own Milky Way Galaxy, also
in physical contact with our galactic disk.
Another small satellite galaxy of our Milky Way Galaxy is the Large Magellanic Cloud, which astronomers always abbreviate the LMC. The Large Magellanic Cloud is roughly fifty kiloparsecs
(roughly one hundred and fifty thousand light-years) from our Solar
System. The famous supernova SN1987A occurred in the Large Magellanic
Cloud, as we discussed earlier in the course.
Yet another small satellite galaxy of our Milky Way Galaxy is the Small Magellanic Cloud, which astronomers always abbreviate the
SMC. The Small Magellanic
Cloud is roughly sixty kiloparsecs (roughly two
hundred thousand light-years) from our Solar System. The two Magellanic
Clouds are visible from the southern hemisphere without the aid of a telescope
or even a pair of binoculars. To the
naked eye, the two Magellanic Clouds appear to be
colorful clouds several times larger than the Full Moon in the sky. The two Magellanic
Clouds are named for the Portuguese explorer Ferdinand
Magellan who led the first mission to successfully circumnavigate the entire
world in the early sixteenth century (the early 1500s). The two Magellanic
Clouds are classified as irregular galaxies. Other small satellite galaxies of our Milky
Way Galaxy include the Draco Dwarf Galaxy, the Sculptor Dwarf Galaxy, the
Carina Dwarf Galaxy, the Fornax Dwarf Galaxy, and the Phoenix Dwarf
Galaxy. Each small satellite galaxy of
our Milky Way Galaxy is composed of roughly one billion stars. As these small satellite galaxies move along
their orbits, the gravity of our Milky Way Galaxy perturbs
the motion of the stars within these small satellite galaxies. As a result, our Milky Way Galaxy slowly rips
apart these small satellite galaxies.
Ultimately, our Milky Way Galaxy will devour some of these small
satellite galaxies, causing our Milky Way Galaxy to gradually
grow larger and larger over billions of years. In fact, there is evidence that some groups
of stars within our Milky Way Galaxy were formerly small satellite galaxies
that our Milky Way Galaxy has completely devoured.
The nearest major galaxy to
our Milky Way Galaxy is the Andromeda Galaxy, roughly eight hundred kiloparsecs (roughly 2.5 million light-years) distant. The Andromeda Galaxy is a spiral-disk galaxy
similar in size, mass, and structure (shape) to our own Milky Way Galaxy. The Andromeda Galaxy also has its own small satellite
galaxies, the two most prominent being M32 (also
designated NGC 221) and M110 (also
designated NGC 205). Other small
satellite galaxies of the Andromeda Galaxy include NGC 185 and NGC 147. The uppercase (capital) letter M refers to
the Messier deep sky catalogue, a list of one hundred and ten faint objects in
the sky compiled by the French astronomer Charles Messier in the eighteenth
century (the 1700s).
The objects on the Messier deep sky catalogue that we
have discussed include M1 the Crab Nebula (a
supernova remnant), M7 the Ptolemy Cluster (an open
star cluster), M13 the Hercules Cluster (a globular
star cluster), M20 the Trifid
Nebula (a diffuse nebula), M31 the Andromeda Galaxy
(a spiral-disk galaxy), M32 (a small satellite galaxy
of M31), M33 the Triangulum
Galaxy (a spiral-disk galaxy that we will discuss shortly), M42
the Orion Nebula (a diffuse nebula), M45 the Pleiades
Cluster (an open star cluster), M57 the Ring Nebula
(a planetary nebula), and M110 (a small satellite
galaxy of M31). The New General Catalogue, which astronomers
always abbreviate NGC, is a more comprehensive deep sky catalogue compiled in
the late nineteenth century (the late 1800s). The Index Catalogue, which astronomers always
abbreviate IC, is an even more comprehensive deep sky catalogue than the New
General Catalogue. Caution: a particular
deep sky object may have one Messier Catalogue number, a different New General
Catalogue number, and yet another Index Catalogue number! A similar confusion occurs with stellar
designations. Any particular star may
have one CPD number (from the Cape Photographic Durchmusterung Catalogue), another GSC
number (from the Guide Star Catalog), yet another HD number (from the Henry
Draper Catalogue), yet another SAO number (from the Smithsonian Astrophysical
Observatory Catalog), and yet another HIP number (from the Hipparcos
Catalogue)! The light from the Andromeda
Galaxy is blueshifted, revealing that the Andromeda
Galaxy is moving toward our Milky Way Galaxy.
In actuality, our Milky Way Galaxy and the Andromeda Galaxy are falling
toward each other due to their mutual gravitational attraction. These two galaxies will collide in roughly
five billion years. We will discuss
galactic collisions in detail shortly.
The Triangulum Galaxy is another nearby major spiral-disk galaxy,
although it is significantly smaller than our Milky Way Galaxy or the Andromeda
Galaxy; the Triangulum Galaxy is composed of roughly ten billion stars. The Triangulum Galaxy is more than eight
hundred kiloparsecs (almost three million
light-years) distant. The Local Galactic
Group is the collection of several dozen galaxies that together define our
galactic neighborhood. Most of the galaxies
in the Local Galactic Group are small irregular galaxies, such as the two Magellanic Clouds.
There are only three major galaxies in the Local Galactic Group: our own
Milky Way Galaxy, the Andromeda Galaxy, and the Triangulum Galaxy. The Local Galactic Group is roughly three megaparsecs (roughly ten million light-years) in diameter.
We cannot measure the
distances to our satellite galaxies such as the two Magellanic
Clouds using the main sequence fitting method, and measuring the distance to
the Andromeda Galaxy or the Triangulum Galaxy is out of the question using this
main sequence fitting method. Therefore,
we need a higher rung of the Cosmological Distance Ladder to measure these
extragalactic distances. The next major
rung of the Cosmological Distance Ladder above the main sequence fitting method
is the variable star method. A variable
star has a luminosity (absolute magnitude or intrinsic brightness) that varies
significantly. These significant
variations are caused by pulsations within the star;
as a variable star expands and contracts, its surface area changes, thus
varying its luminosity. There are many
different classes of variable stars, such as Cepheid variable stars, Lyrae variable stars, Mira variable stars, and Tauri variable stars.
As we discussed earlier in the course, Tauri
variable stars are protostars, Cepheid variable stars
are transitioning from the main sequence along the first asymptotic giant
branch, Lyrae variable stars are horizontal-branch
stars, and Mira variable stars are transitioning from the helium-burning phase
along the second asymptotic giant branch.
All variable stars are transitioning from one equilibrium evolutionary
stage to another equilibrium evolutionary stage. A transition is essentially an instability,
thus causing pulsations within the star, causing its size to oscillate from
large to small and back again. As a
result, the luminosity of a variable star oscillates from bright to dim and
back again. At the beginning of the
twentieth century (the early 1900s), the American
astronomer Henrietta Leavitt discovered an equation relating the average
luminosity of Cepheid variable stars with their pulsation period. This equation is called
the Leavitt period-luminosity relation in her honor. Other similar equations have
been discovered for other classes of variable stars, and all such
equations are known as period-luminosity relations. We can use these period-luminosity relations
to measure extragalactic distances.
First, we determine the distance to nearby variable stars within our own
Milky Way Galaxy using the parallax method or the main sequence fitting
method. We combine their distance with
their average apparent magnitude to calculate their average absolute
magnitude. We also measure the pulsation
period of these nearby variable stars; their pulsation period together with their
average luminosity establishes the period-luminosity relations. Now suppose we discover variable stars within
another galaxy. Even nearby galaxies are
sufficiently distant that we cannot use the parallax method or the main
sequence fitting method to measure their distance. Instead, we measure the pulsation periods of
the variable stars we have discovered within these galaxies. Using the established period-luminosity relations,
we can now calculate the average luminosity of these variable stars. Finally, we combine their average luminosity
with their average apparent magnitude to determine the distance to these
variable stars and hence the distance to the galaxy wherein they reside. This procedure is called
the variable star method, and it is the next major rung of the Cosmological
Distance Ladder above the main sequence fitting method. To determine the distance to a nearby galaxy,
we measure the pulsation periods of variable stars within the galaxy. We then use the established period-luminosity
relations to determine the average luminosity of these variable stars, and
finally we combine the average luminosity with the average apparent magnitude
to calculate the distance. This variable
star method is used to determine distances to galaxies
throughout the Local Galactic Group. In
fact, this variable star method is used to measure
distances to galaxies even beyond the Local Galactic Group, out to distances of
a couple hundred megaparsecs (a few hundred million
light-years) from our Milky Way Galaxy.
During the eighteenth century
(the 1700s), moderately powerful telescopes began to
reveal faint irregular objects in the sky that astronomers called nebulae. Somewhat more powerful
telescopes in the nineteenth century (the 1800s)
magnified these nebulae. Some of these
nebulae still appeared irregularly shaped when magnified, but other nebulae
appeared to have spiral shapes when magnified.
Astronomers named these objects spiral nebulae. At the beginning of the twentieth century
(the early 1900s), some astronomers claimed that
these spiral nebulae were not nebulae at all.
These astronomers claimed that these objects were actually collections
of billions of stars held together by gravity.
In other words, these astronomers claimed that the universe is not homogeneously filled with stars; these astronomers
claimed that stars are clumped into gigantic organizations that they called
island universes. These astronomers also
claimed that we live within one of these island universes as revealed by the
band of milk that wraps around our sky.
Other astronomers were opposed to this new idea; these astronomers
claimed that spiral nebulae were clouds of gas and nothing more. The debate over the true nature of spiral
nebulae is called the Great Debate in the history of
astronomy. This Great Debate occurred on
April 26, 1920, at the Smithsonian Museum of Natural History in Washington,
D.C., between the two American astronomers Harlow Shapley and Heber Curtis. Therefore, the Great Debate is also called the Shapley-Curtis Debate. Curtis argued in favor of island universes,
while Shapley argued against island universes.
Neither of these astronomers settled this Great Debate. The greatest American astronomer of the
twentieth century, Edwin Hubble, settled this Great Debate a few years
later. In the year 1924, Edwin Hubble
discovered Cepheid variable stars in what was then called
the Andromeda Spiral Nebula and the Triangulum Spiral Nebula. Using the Leavitt period-luminosity relation,
Edwin Hubble calculated the distances to these so-called spiral nebulae to be
in the hundreds of kiloparsecs (millions of
light-years). The only way we could ever
see anything at such incredible distances is if it shines with the luminosity
of billions of stars. Therefore,
astronomers realized that Heber Curtis was correct. These so-called spiral nebulae are not
nebulae at all; they are island universes of billions of stars, and we live
within one of these island universes.
Eventually, these island universes were renamed galaxies. The word galaxy is derived
from the Greek root galacto- for milk. For example, galactose and glucose are the
two simple sugars that together compose the milk sugar lactose. The Andromeda Spiral Nebula was renamed the
Andromeda Galaxy, the Triangulum Spiral Nebula was renamed the Triangulum
Galaxy, and our home galaxy was named the Milky Way Galaxy, which literally
means milky milk!
As we observe galaxies beyond
the Local Galactic Group, we discover that there are two main types of galaxies
in our universe. One type is spiral-disk
galaxies, such as our own Milky Way Galaxy, the Andromeda Galaxy, and the
Triangulum Galaxy. The other major type
of galaxy in our universe is elliptical galaxies. These elliptical galaxies should
really be called ellipsoidal galaxies, since their true shape is a
three-dimensional ellipse, and a three-dimensional ellipse is called an
ellipsoid. Nevertheless, astronomers
named them elliptical galaxies, since their shapes appear to be ellipses in
photographs. Spiral-disk galaxies are
more flat in structure (shape), while elliptical galaxies are more round in
structure (shape). Spiral-disk galaxies
are more flat because most of their stars orbit their galactic center in nearly
the same plane in nearly the same angular (orbital) direction. The orbits of all of these stars add together
to give spiral-disk galaxies high angular momentum. Elliptical galaxies are more round because
most of their stars orbit their galactic center in random orbits in random
directions. The orbits of all these
stars mostly cancel each other to give elliptical galaxies low angular
momentum. Spiral-disk galaxies have an
abundance of interstellar gas resulting in active star formation, making their
stellar populations relatively high mass, hot, luminous, and blue (early-type
stars) with high metallicities.
Elliptical galaxies have little interstellar gas and hence little star
formation, making their stellar populations relatively low mass, cool, dim, and
red (late-type stars) with low metallicities.
Of course, when an elliptical galaxy was first born, some of its stars
must have been high mass, hot, luminous, and blue (early-type stars). However, these stars have short main-sequence
lifetimes, as we discussed earlier in the course. With very little interstellar gas to give
birth to new stars, the only stars remaining in an elliptical galaxy after the
short lifetimes of the early-type stars are low mass, cool, dim, and red
(late-type stars) with low metallicities.
The stellar populations within spiral-disk galaxies and elliptical
galaxies should sound familiar. Within our Milky Way Galaxy, the relatively high mass, hot,
luminous, and blue (early-type) stars orbiting in roughly the same angular
(orbital) direction within the galactic disk where an abundance of interstellar
gas results in active star formation are the Population I stars, while the
relatively low mass, cool, dim, and red (late-type) stars orbiting in random
directions throughout the galactic halo where there is little interstellar gas
and hence little star formation are the Population II stars. We conclude that a spiral-disk galaxy is an
entire galaxy of mostly Population I stars, while an elliptical galaxy is an
entire galaxy of mostly Population II stars.
Spiral-disk galaxies have
high angular momentum due to most of its stars orbiting its galactic center in
nearly the same plane in nearly the same angular (orbital) direction. As a result, the light we receive from one
side of a spiral-disk galaxy is blueshifted relative
to its galactic center, since those stars happen to be moving toward us, while
the light we receive from the other side of the spiral-disk galaxy is
redshifted relative to its galactic center, since those stars happen to be
moving away from us. From these blueshifts and redshifts, we can calculate the speeds with
which the stars orbit the galactic center.
From these speeds, we can calculate the gravitational force acting on
these stars, and hence we can calculate the distribution of mass within these
galaxies. Again, we discover that there
is roughly ten times as much mass as normal (luminous star) mass. This is the mysterious dark matter. Although elliptical galaxies have low angular
momentum, we can still measure a dispersion of blueshifts
and redshifts in the light from these galaxies, enabling us to calculate a
velocity dispersion. Again, we can
calculate the gravitational force acting on these stars to cause the velocity
dispersion, and again we can calculate the distribution of mass within these
galaxies. Yet again, we discover that
there is roughly ten times as much dark matter as normal (luminous star)
matter. Evidently, all galaxies in the
universe are composed of roughly ninety percent dark matter and only roughly
ten percent normal (luminous star) matter.
All of these observations provide us with another method of determining
distance. The Tully-Fisher relation is
an equation that correlates the orbital speed of stars within a spiral-disk
galaxy to the luminosity of the spiral-disk galaxy, named for astronomers R.
Brent Tully and J. Richard Fisher who together first formulated this
equation. The orbital speed of the stars
within a spiral-disk galaxy is caused by the
gravitational force, which is exerted by the total mass of the galaxy. Although the total mass of the galaxy is
mostly dark matter, there is still a correlation between the total amount of
mass and the amount of normal (luminous star) mass. If blueshifts and
redshifts are more severe, the stars must be orbiting faster from a stronger
gravitational force caused by a greater quantity of total mass, both dark
matter and normal (luminous star) matter.
If blueshifts and redshifts are more modest,
the stars must be orbiting slower from a weaker gravitational force caused by a
lesser quantity of total mass, both dark matter and normal (luminous star)
matter. In brief, the Tully-Fisher
relation states that if a spiral-disk galaxy rotates faster, then it must be
more luminous, and if a spiral-disk galaxy rotates slower, then it must be less
luminous. The Faber-Jackson relation is
a similar equation that correlates the velocity dispersion of stars within an
elliptical galaxy to the luminosity of the elliptical galaxy, named for
astronomers Sandra Faber and Robert Jackson who together first formulated this
equation. To use the Tully-Fisher
relation and the Faber-Jackson relation to determine the distance to distant
galaxies, we first use the variable star method to determine the distance to
somewhat closer galaxies. We combine the
distance with the apparent magnitude of these somewhat closer galaxies to
calculate the luminosity or the absolute magnitude or the intrinsic brightness
of these somewhat closer galaxies. Some
of these somewhat closer galaxies are spiral-disk galaxies, while others are
elliptical galaxies. We measure the
orbital speed of stars within the spiral-disk galaxies to establish the
Tully-Fisher relation, and we measure the velocity dispersion of stars within
the elliptical galaxies to establish the Faber-Jackson relation. Now suppose we wish to measure the distance
to galaxies so distant that we cannot use the variable star method, since even
our most powerful telescopes cannot resolve variable stars within these remote
galaxies. For distant spiral-disk
galaxies, we measure the orbital speed of its stars, and we use the established
Tully-Fisher relation to calculate the luminosity of the spiral-disk
galaxy. For distant elliptical galaxies,
we measure the velocity dispersion of its stars, and we use the established
Faber-Jackson relation to calculate the luminosity of the elliptical
galaxy. In either case, we combine the
luminosity with the apparent magnitude to finally calculate
the distance to the galaxy. The
Tully-Fisher relation and the Faber-Jackson relation together are the next
major rung (above the variable star method) of the Cosmological Distance
Ladder.
Not only are stars clumped
together to form galaxies, but galaxies are themselves clumped together to form
even larger organizations called galactic groups or galactic clusters. Galactic groups are composed of several dozen
galaxies, although most of these galaxies are small minor galaxies. A typical galactic group is composed of less
than ten large major galaxies. For
example, our Local Galactic Group is composed of several dozen small minor
galaxies but only three large major galaxies: our own Milky Way Galaxy, the
Andromeda Galaxy, and the Triangulum Galaxy.
The nearest galactic group to our own Local Galactic Group is the Maffei Galactic Group, roughly three megaparsecs
(roughly ten million light-years) distant.
Other nearby galactic groups include the Bode Galactic Group at roughly
3.5 megaparsecs (more than eleven million
light-years) distant, the Sculptor Galactic Group at nearly four megaparsecs (nearly thirteen million light-years) distant,
and the Leo Triplet Galactic Group at nearly eleven megaparsecs
(roughly thirty-five million light-years) distant. Each of these galactic groups is composed of
only a few large major galaxies and several dozen small minor galaxies. Galactic clusters on the other hand are
composed of hundreds of large major galaxies.
The nearest galactic cluster is the Virgo Galactic Cluster, roughly seventeen
megaparsecs (more than fifty-six million light-years)
distant. Other nearby
galactic clusters include the Fornax Galactic Cluster at nearly twenty megaparsecs (more than sixty million light-years) distant,
the Eridanus Galactic Cluster at roughly twenty-three megaparsecs
(roughly seventy-five million light-years) distant, the Antila
Galactic Cluster at more than forty megaparsecs (more
than 130 million light-years) distant, the Centaurus Galactic Cluster at more
than fifty megaparsecs (more than 170 million
light-years) distant, and the Hydra Galactic Cluster at nearly sixty megaparsecs (nearly two hundred million light-years)
distant. Each of these galactic
clusters is composed of hundreds of large major galaxies. We can determine the distance to a galactic
cluster by applying the Tully-Fisher relation to its spiral-disk members and
the Faber-Jackson relation to its elliptical members. The results using these two different methods
on galaxies within the same galactic cluster have always been
found to be roughly consistent with each other, confirming the
reliability of these two methods for determining distance. We can also measure the blueshifts
and the redshifts of entire galaxies within a galactic cluster relative to the
center of the galactic cluster to determine the orbital speeds of entire
galaxies within the galactic cluster. We
can then calculate the gravitational force responsible for these orbital
speeds, and thus we can calculate the total mass that exerts this gravitational
force. We discover that galactic
clusters are composed of roughly ten times as much dark matter as normal
(luminous star) matter. There is also
diffuse gas that fills the space between galaxies within galactic
clusters. This diffuse gas is called the intergalactic medium, since this gas is
distributed among many galaxies. This
diffuse gas is also called the intracluster
medium, since this gas is distributed within a galactic cluster. Caution: the prefix inter- means among, while
the prefix intra- means within. Recall
that the gas that fills the space among the stars of the galactic disk of our
Milky Way Galaxy is called the interstellar medium,
since this gas is distributed among billions of stars. Note however that this interstellar medium could also be called the intragalactic
medium, since this gas is within our Milky Way Galaxy. We have measured the temperature of the
intergalactic/intracluster medium within galactic
clusters to be in the millions of kelvins, since this gas radiates primarily
X-rays. Note therefore that the
intergalactic/intracluster medium within galactic
clusters was not discovered until astronomers placed
X-ray telescopes in orbit around the Earth, since the Earth’s atmosphere is
opaque to X-rays, as we discussed earlier in the course. We can calculate the gravitational force
necessary to heat the intergalactic/intracluster
medium to these extremely hot temperatures, and again we can calculate the
total mass necessary to exert this gravitational force. Yet again, we discover that galactic clusters
are composed of roughly ten times as much dark matter as normal (luminous star)
matter. A galactic cluster contains so
much mass that the gravity of a galactic cluster acts as a gravitational lens,
bending the light from even more distant galaxies from behind the cluster along
our line of sight, as we discussed earlier in the course. We can use the curved images of these distant
galaxies to calculate the gravitational force that causes this lensing, and yet
again we can calculate the total mass necessary to exert this gravitational
force. Yet again, we discover that galactic
clusters are composed of roughly ten times as much dark matter as normal
(luminous star) matter. Whenever we use these three different methods to calculate the
total mass of any particular galactic cluster (the orbital speeds of entire
galaxies within the galactic cluster, the temperature of the intergalactic/intracluster medium within the galactic cluster, and the
gravitational lensing of distant galaxies caused by the galactic cluster), the
results have always been found to be roughly consistent with one another. Not only does this
consistency confirm the reliability of these three methods of determining the
total mass of galactic clusters, but this consistency together with rotation
curves of individual spiral-disk galaxies (the Tully-Fisher relation) and
velocity dispersions of individual elliptical galaxies (the Faber-Jackson
relation) all provide strong evidence that the entire universe is composed of
roughly ten times as much dark matter as normal (luminous star) matter.
Not only are stars clumped
together to form galaxies and not only are galaxies themselves clumped together
to form galactic groups or galactic clusters, but galactic groups and galactic
clusters are clumped into enormous organizations called galactic superclusters. Our own Local Galactic
Group, the Maffei Galactic Group, the Bode Galactic
Group, the Sculptor Galactic Group, the Leo Triplet Galactic Group, and
innumerable other galactic groups as well as the Virgo Galactic Cluster, the
Fornax Galactic Cluster, the Eridanus Galactic Cluster, the Antila
Galactic Cluster, the Centaurus Galactic Cluster, the Hydra Galactic Cluster,
and several other galactic clusters all together form the Laniakea
Galactic Supercluster. The Laniakea Galactic Supercluster is more than 150 megaparsecs (more than five hundred million light-years) in
diameter, and our own Local Galactic Group as well as the nearby Maffei Galactic Group, the Bode Galactic Group, the
Sculptor Galactic Group, and the Leo Triplet Galactic Group are all on the
outskirts of the Laniakea Galactic Supercluster. Nearby galactic
superclusters to our own Laniakea Galactic
Supercluster include the Hydra-Centaurus Galactic Supercluster at roughly
seventy megaparsecs (more than 200 million
light-years) distant, the Perseus-Pisces Galactic Supercluster at nearly eighty
megaparsecs (roughly 250 million light-years)
distant, the Coma Galactic Supercluster at more than ninety megaparsecs
(roughly three hundred million light-years) distant, and the Shapley Galactic
Supercluster at roughly two hundred megaparsecs
(roughly 650 million light-years) distant. Each of these galactic superclusters contains
many galactic groups and several galactic clusters. There are enormous regions between galactic
superclusters that are nearly empty of galaxies. These enormous regions are
called cosmic voids. Moreover,
galactic superclusters are clumped into colossal
organizations called cosmic filaments, each containing between a few hundred
thousand and a few million galaxies. Our
own Laniakea Galactic Supercluster, the
Hydra-Centaurus Galactic Supercluster, the Perseus-Pisces Galactic
Supercluster, and several other galactic superclusters all together form the
Perseus-Pisces-Sculptor-Hercules Cosmic Filament. Nearby cosmic filaments to our own Perseus-Pisces-Sculptor-Hercules
Cosmic Filament include the Coma-Hercules-Leo Cosmic
Filament, the Sculptor Cosmic Filament, and the Sloan Cosmic Filament. There are colossal regions between cosmic
filaments that are nearly completely empty of galaxies. These colossal regions are
called cosmic supervoids. At size scales of hundreds
of megaparsecs (hundreds of millions of light-years)
and smaller, the universe appears rather clumpy, with stars clumped together to
form galaxies, galaxies clumped together to form galactic groups or galactic
clusters, galactic groups and galactic clusters clumped together to form
galactic superclusters with cosmic voids between them, and galactic
superclusters clumped together to form cosmic filaments with cosmic supervoids between them. However, at size scales of gigaparsecs (billions of light-years), the universe appears
less clumpy and more homogeneous (more smooth), since many cosmic filaments
appear close to each other relative to these titanic size scales. In other words, there appear to be no
conglomerations of matter larger than cosmic filaments. Cosmologists call this homogeneous (smooth)
distribution of mass at these titanic size scales the end of greatness. The observable universe contains hundreds of
thousands, perhaps millions, of cosmic filaments. Since each cosmic filament contains between a
few hundred thousand and a few million galaxies, the entire observable universe
contains roughly one hundred billion galaxies.
Assuming that each galaxy contains on average one hundred billion star
systems just like our Milky Way Galaxy, then there are
roughly ten sextillion star systems in the observable universe.
To measure these incredible
distances, we need an even higher rung of the Cosmological Distance
Ladder. As we discussed earlier in the
course, high mass stars end their lives with Type II supernova explosions. Consider instead a white dwarf orbiting a
giant star in a close binary star system with a mass transfer from the giant star
to the white dwarf through an accretion disk around the white dwarf, as we
discussed earlier in the course. This
white dwarf will suffer from a nova once every few decades, once every few
centuries, or once every few millennia, periodically ejecting material from its
surface triggered by the fusion of hydrogen into helium on its surface, as we
also discussed earlier in the course.
These periodic novae do not stop the mass transfer from the giant star
to the white dwarf from continuing.
Hence, the mass of the white dwarf will change. This variation in mass of the white dwarf
depends upon the mass transfer rate from the giant star as
compared with the mass loss rate from the periodic novae. If the mass loss rate from the periodic novae
happens to be greater than the mass transfer rate from the giant star, then the
white dwarf will lose more and more mass over time. If the mass transfer rate from the giant star
happens to be greater than the mass loss rate from the periodic novae, then the
white dwarf will gain more and more mass over time, but the increasing mass of
the white dwarf cannot continue indefinitely.
The maximum mass of a white dwarf is the Chandrasekhar limit, equal to 1.4M☉ (1.4
solar masses), as we discussed earlier in the course. If a white dwarf gains so much mass that it reaches
this Chandrasekhar limit, electron degeneracy pressure will no longer be able
to support the white dwarf. The
self-gravity of the white dwarf will force the white dwarf to compress, raising
the temperature of the white dwarf until carbon fusion is initiated. White dwarfs are composed almost entirely of
carbon, as we discussed earlier in the course.
Thus, the entire white dwarf suffers from a thermonuclear detonation,
obliterating the entire white dwarf in a cataclysmic explosion that liberates
energy in the billions of solar luminosities!
This is roughly the total power output of an entire galaxy of
stars! This unique type of explosion is called a Type Ia supernova, as
opposed to a Type II supernova.
Observationally, the Type II supernova of a high mass star and the Type Ia supernova of a carbon white dwarf may seem identical, at
least at first glance. However, a Type
II supernova occurs when the iron-nickel white dwarf core of a red supergiant can no longer be supported by electron degeneracy pressure. The photons liberated from the collapse of
the iron-nickel white dwarf core must pass through the outer layers of the red
supergiant, which are composed of mostly hydrogen. Therefore, the light we receive from a Type
II supernova has strong hydrogen absorption lines. With a Type Ia
supernova however, the carbon white dwarf is mostly naked, aside from the
surrounding accretion disk and the nearby giant star that it orbits. Thus, the photons liberated from a Type Ia supernova pass through hardly any outer gas layers;
therefore, the light we receive from a Type Ia
supernova has weak hydrogen absorption lines.
This is one way astronomers discriminate between a
Type Ia supernova and a Type II supernova. The light we receive from a Type II supernova
that results from the collapse of the iron-nickel white dwarf core of a red
supergiant has strong hydrogen absorption lines, while the light we receive
from a Type Ia supernova that results from the
thermonuclear detonation of a mostly naked carbon white dwarf has weak hydrogen
absorption lines. Astronomers also use
light curves to discriminate between Type Ia
supernovae and Type II supernovae. A
light curve is a graph of the amount of light we receive from anything in the
universe plotted as a function of time.
The light curves from Type Ia
supernovae and Type II supernovae have different shapes. In particular, the light curve of a Type Ia supernova has a more steep
decline, while the light curve of a Type II supernova has a more gradual
decline. Notice that all Type Ia supernovae occur in exactly the
same way; a carbon white dwarf slowly gains mass until it reaches the
Chandrasekhar limit, which has a very specific value. Therefore, all Type Ia supernovae have the same luminosity. Notice that all Type II
supernovae do not occur in exactly the same way. There is a wide range of masses for high mass
stars, from 7M☉, 8M☉, or 9M☉ (seven,
eight, or nine solar masses) all the way up to the Eddington limit of roughly 100M☉ (one
hundred solar masses), that may suffer from a Type II supernova. Therefore, Type II supernovae have a wide
range in luminosities. We conclude that
Type Ia supernovae are standard candles, while Type
II supernovae are not standard candles. The term standard candle was used by astronomers more than
one hundred years ago when humans still used candles to light their homes! Perhaps a better term today would be standard
lightbulb instead of standard candle.
Nevertheless, astronomers continue to use the term standard candle. Imagine a candle or a lightbulb we see in the
distance. We cannot determine the
distance to that light source unless we knew its luminosity (its power
output). If we knew the luminosity (the
power output) of the light source, we could easily combine the luminosity of
the light source with the apparent brightness of the light source to calculate
its distance from us. This is what we
mean by the term standard candle, a light source with a known luminosity (power
output) that we may combine with its apparent brightness to determine its
distance. Type Ia supernovae are standard candles, since they all
have the same luminosity. Type II
supernovae are not standard candles, since they have a wide range in
luminosities. We can therefore use Type Ia supernovae to determine the
distance to extremely remote galaxies.
We begin by observing Type Ia
supernovae within somewhat closer galaxies.
We can use the variable star method or the
Tully-Fisher relation or the Faber-Jackson relation to determine the
distance to these somewhat closer galaxies.
We then combine the distance with the apparent brightness of the Type Ia supernovae to determine the luminosity or the absolute
magnitude or the intrinsic brightness of Type Ia supernovae.
As we discussed earlier in the course, astronomers continuously monitor
tens of thousands of galaxies and thus observe hundreds of distant supernovae
every year. If we happen to discover a
supernova in an extremely remote galaxy with strong hydrogen absorption lines
with a gradual decline in its light curve, then we are out of luck, since this
is a Type II supernova, which is not a standard candle. If instead we happen to discover a supernova
in an extremely remote galaxy with weak hydrogen absorption lines with a steep
decline in its light curve, then this must be a Type Ia
supernova, which is a standard candle.
We combine its apparent brightness with the absolute magnitude of all
Type Ia supernovae to
determine the distance to that supernova and hence the distance to the galaxy
wherein it resides. This is called the Type Ia supernova
method, and it is one of the highest rungs of the Cosmological Distance Ladder,
since supernovae are so incredibly luminous that we can observe them
practically across the entire observable universe.
The Hubble Classification of Galaxies
Edwin Hubble, the greatest
American astronomer of the twentieth century, classified galaxies based on
their structural appearance (their shape).
He designated elliptical galaxies that appeared perfectly round as E0. He designated
elliptical galaxies that appeared almost perfectly round but slightly elongated
as E1. The
next designation is E2 for elliptical galaxies
appearing even less round and hence even more elongated. The next designation is E3
for elliptical galaxies appearing moderately elongated. Hubble designated elliptical galaxies that
appeared even more elongated as E4. The E5 elliptical
galaxies appear quite elongated, and E6 elliptical
galaxies appear even more elongated.
Hubble designated the most elongated elliptical galaxies as E7. Edwin Hubble
then grouped spiral-disk galaxies into two subcategories: unbarred spirals and
barred spirals. A barred spiral galaxy
happens to have billions of its stars lined up along the shape of a rod or a
bar through its central bulge; an unbarred spiral galaxy does not have a bar
through its central bulge. As we
discussed, the structure (the shape) of a galaxy is
determined by the orbits of all of its individual stars. Hence, the structure (the shape) of a galaxy
continuously changes over millions of years as all of its
stars move along their orbits. The
spiral arms of a spiral-disk galaxy are not permanent structures, as we
discussed. Similarly, central bars of
spiral-disk galaxies are not permanent structures. A barred spiral-disk galaxy may have been an
unbarred spiral-disk galaxy millions of years ago, and it may become an
unbarred spiral-disk galaxy again millions of years from now. Similarly, an unbarred spiral-disk galaxy may
have been a barred spiral-disk galaxy millions of years ago, and it may become
a barred spiral-disk galaxy again millions of years from now. Edwin Hubble also classified spiral-disk
galaxies (both unbarred and barred) based on the size of their central bulge,
the wrapping of their spiral arms, and the smoothness versus the clumpiness of their spiral arms. The clumpy regions within the arms of some
spiral-disk galaxies are star clusters and diffuse nebulae within their
interstellar gases. For unbarred spiral-disk
galaxies, beginning with unbarred spirals with a large central bulge and tightly-wrapped, smooth arms, the galaxy types are Sa, Sab,
Sb, Sbc, and finally Sc for
the unbarred spirals with a small central bulge and loosely-wrapped, clumpy
arms. Barred spiral-disk galaxies have a
similar classification, except that Hubble added an uppercase (capital) letter
B meaning barred. Beginning with barred
spirals with a large central bulge and tightly-wrapped,
smooth arms, the galaxy types are SBa, SBab, SBb, SBbc,
and finally SBc for the barred spirals with a small
central bulge and loosely-wrapped, clumpy arms.
We would read Sb for example as “unbarred spiral b,” and we would read SBb for example as “barred spiral b.” Edwin Hubble arranged this classification
scheme into a diagram that resembles a tuning fork, with elliptical galaxies
from E0 through E7 being
the handle of the tuning fork and the spiral galaxies being the two teeth of
the tuning fork. The unbarred spirals
from Sa through Sc form one
tooth of the tuning fork, while the barred spirals from SBa
through SBc form the other tooth of the tuning
fork. Note that Hubble placed irregular
galaxies outside of this tuning fork classification diagram, although he did
give irregular galaxies the designation Irr. Also note that there
is a fourth type of galaxy called lenticular galaxies that have some properties
of spiral-disk galaxies and some properties of elliptical galaxies. In particular, lenticular galaxies have
little interstellar gas resulting in little star formation, making their stellar
populations relatively low mass, cool, dim, and red (late-type stars), like
elliptical galaxies. However, lenticular
galaxies are more flat in structure (shape) because most of their stars orbit
their galactic center in nearly the same plane in nearly the same angular
(orbital) direction, giving lenticular galaxies high angular momentum like
spiral-disk galaxies. The flat structure
(shape) of a lenticular galaxy may also be either unbarred or barred, again
like spiral-disk galaxies. Note however
that the flat structure (shape) of lenticular galaxies does not include strong
spiral arms, unlike spiral-disk galaxies.
For all of these reasons, Hubble placed lenticular galaxies between
elliptical galaxies and spiral-disk galaxies in his classification scheme,
designating the unbarred lenticular galaxies as S0
and the barred lenticular galaxies as SB0. Therefore, a more complete sequence of
unbarred galaxies along one tooth of the Hubble tuning fork is S0, Sa, Sab, Sb, Sbc, and finally Sc.
Similarly, a more complete sequence of barred galaxies along the other
tooth of the Hubble tuning fork is SB0, SBa, SBab,
SBb, SBbc, and finally SBc.
Within this Hubble
classification scheme, our own Milky Way Galaxy is classified
as type SBb, since our Milky Way Galaxy has a bar
through its central bulge, its central bulge is moderate in size, its spiral
arms are moderately wrapped, and its spiral arms are moderately clumpy. In nearly every way imaginable, we live in an
ordinary place in the universe. Firstly,
our planet Earth is not the center of our Solar System; our Sun is at the
center of our Solar System. The Earth is
not the closest planet to the
Sun, nor is the Earth the furthest planet from the Sun. The Earth is the third planet from the Sun,
which is somewhat intermediate in the order of planets
of our Solar System. Our Sun is not a
high mass, hot, luminous star, nor is our Sun a low mass, cool, dim star. Our Sun is intermediate in mass, intermediate
in temperature, and intermediate in luminosity.
Our Sun is not a young star, nor is our Sun an old star. Our Sun is intermediate in age. Our Solar System is not at the center of our
Milky Way Galaxy; a supermassive black hole is at the center of a barred
galactic bulge at the center of our Milky Way Galaxy. Our Solar System is not near the galactic
center, nor is our Solar System far from the galactic center. Our Solar System is roughly halfway out from
the galactic center. Our Milky Way
Galaxy does not have a particularly large central bulge with tightly-wrapped,
smooth spiral arms, nor does our Milky Way Galaxy have a particularly small
central bulge with loosely-wrapped, clumpy arms. Our Milky Way Galaxy has a moderately-sized
central bulge with moderately-wrapped spiral arms that are moderately
clumpy. Again, in nearly every way
imaginable, we live in an ordinary place in the universe.
When Edwin Hubble first
constructed his tuning-fork shaped classification diagram, he believed that
this diagram revealed a galactic evolutionary sequence. In particular, Hubble
believed that supposedly all galaxies are born E0
then supposedly become E1 followed by E2 then E3 then E4 then E5 then E6 then E7 followed supposedly by
lenticular (either unbarred or barred) followed supposedly by a-type
spiral-disk (either unbarred or barred) then ab-type spiral-disk (either
unbarred or barred) then b-type spiral-disk (either unbarred or barred) then bc-type spiral-disk (either unbarred or barred) then c-type
spiral-disk (either unbarred or barred).
Finally, Hubble believed that all galaxies supposedly die as irregular
galaxies. Today, we realize that this
evolutionary sequence is not correct.
Unfortunately, many astronomers believed Edwin Hubble so strongly that
the Hubble classification of galaxies was called the
Hubble sequence, and many astronomers believed that the Hubble sequence was an
evolutionary sequence. Hubble and other
astronomers believed so strongly that the Hubble sequence was an evolutionary
sequence that they called elliptical galaxies early-type galaxies, and they
called spiral-disk galaxies late-type galaxies.
Most unfortunately, this incorrect nomenclature persists among
astronomers and astrophysicists to the present day, even though astronomers and
astrophysicists do recognize that the Hubble sequence is not an evolutionary
sequence. For example, astronomers and
astrophysicists may refer to a b-type spiral-disk galaxy as being earlier than
a c-type spiral-disk galaxy, even though they recognize that the Hubble sequence
is not an evolutionary sequence. As
another example, astronomers and astrophysicists may refer to an E5 elliptical galaxy as being later than an E2 elliptical galaxy, even though
they recognize that the Hubble sequence is not an evolutionary sequence. Since this incorrect nomenclature persists
among astronomers and astrophysicists to the present day, we will also use this
incorrect nomenclature in this course.
Note the extraordinary historical parallel between the Hubble sequence
for galaxies and the main sequence for stars.
Just as Ejnar Hertzsprung
and Henry Norris Russell incorrectly believed that the main sequence is an
evolutionary sequence for stars, Edwin Hubble incorrectly believed that the
Hubble sequence is an evolutionary sequence for galaxies. Just as astrophysicists today continue to incorrectly refer to stars as early-type or late-type
depending on their position along the main sequence, astrophysicists today
continue to incorrectly refer to galaxies as early-type or late-type depending
on their position along the Hubble sequence.
We emphasize again that the Hubble sequence is not an evolutionary
sequence, but if galaxies do not evolve along the Hubble sequence, then how do
galaxies actually evolve? How are galaxies actually born? How do galaxies actually live? How do galaxies actually die?
Galactic Evolution: Birth, Life, and Death
Unfortunately, galactic
evolution is not well understood. This is forgivable, since a galaxy is a
collection of billions of star systems.
In other words, a single galaxy is an incredibly complicated dynamical
system. Presumably, a galaxy is born
from a protogalactic cloud, an enormous cloud of gas
millions of light-years across, and presumably a galaxy forms as the protogalactic cloud collapses from its own self-gravity
(under its own weight), but precisely how this galactic formation occurs is not
well understood. There are two main
competing theories to explain galactic birth: the density-angular-momentum
theory and the collision-merger theory.
According to the
density-angular-momentum theory of galactic birth, a galaxy is born somewhere
along the Hubble sequence. This theory
sounds similar to the manner in which stars are born. A star is born somewhere along the main
sequence, as we discussed earlier in the course. A star is born with a particular spectral
type along the main sequence depending on its mass. If a star happens to be born with high mass,
it is born early on the main sequence, while if a star happens to be born with
low mass, it is born late on the main sequence, as we discussed earlier in the
course. According to the
density-angular-momentum theory of galactic birth, a galaxy is born with a
particular Hubble type along the Hubble sequence depending on the density and
the angular momentum of the protogalactic cloud from
which it formed. In particular, if the protogalactic cloud happened to have high density and low
angular momentum, then the protogalactic cloud will
be born an elliptical galaxy, early on the Hubble sequence. If the protogalactic
cloud happened to have low density and high angular momentum, then the protogalactic cloud will be born a spiral-disk galaxy, late
on the Hubble sequence. The detailed
arguments of this density-angular-momentum theory are as follows. If the protogalactic
cloud happened to have low angular momentum, this means that its gases were
moving along random trajectories in random directions. All these orbits mostly cancel each other,
giving the protogalactic cloud low angular
momentum. If in addition the protogalactic cloud happened to have high density, then
stars will form from this high-density gas and continue moving along these
random orbits. If the stars that make up
the resulting galaxy move along mostly random orbits, then the overall shape of
the resulting galaxy will be more round and less flat. Also, there will be
very little gas remaining to form additional stars, since the high-density gas
was mostly consumed to create the stars in the first place. The result is a round-shaped galaxy with
stars moving along random orbits with very little gas to form new stars, but
this is an elliptical galaxy. If on the
other hand the protogalactic cloud happened to have
low density, then many stars will not form yet.
We must wait until the protogalactic cloud
collapses further for the gases to attain a high enough density to form
stars. If in addition the protogalactic cloud happened to have high angular momentum,
then the protogalactic cloud will collapse into a
flat, rotating disk perpendicular to its axis of total angular momentum. More precisely, small higher density regions
within the protogalactic cloud will collide more
frequently as the protogalactic cloud collapses,
since the gravitational collapse brings these regions closer together. Several higher density regions often merge
into larger regions as a result of these
collisions. By the law of conservation
of translational (linear) momentum, the resulting larger regions will have less
motion along the direction of the axis defined by the total angular momentum of
the collapsing protogalactic cloud, since the
collisions will average out their more random motions in this direction. By the law of conservation of angular
momentum, the resulting larger regions will have more
circular orbits, since the collisions will average out their more random
orbits, many of which were more elliptical.
In summary, the laws of physics together cause the gravitationally
collapsing protogalactic cloud to flatten into a
circular, rotating disk perpendicular to the axis of the total angular momentum
of the forming galaxy. The result is a
flat-shaped galaxy with stars all orbiting the galactic center in roughly the
same angular (orbital) direction, but this is a spiral-disk galaxy. Note that a small number of stars would have
formed from sufficiently dense regions even before the high-angular momentum protogalactic cloud collapses. These would be the first stars born within
the resulting spiral-disk galaxy. These
particular stars have lower metallicities, since they were born first when the
universe was younger and hence there was less time for earlier generations of
high mass stars to synthesize metals.
These particular stars would also remain on their original more random
orbits around the resulting spiral-disk galaxy, since they formed before the
high-angular momentum protogalactic cloud collapsed
into a circular, rotating disk. This
small number of stars are the Population II stars within the galactic halo of
the resulting spiral-disk galaxy. Since
the protogalactic cloud happened to have low density
initially, only a small number of Population II stars would be born, and hence
most of the stars are born later after the collapse of the protogalactic
cloud. This large number of stars that
are born later are the Population I stars within the galactic disk of the
resulting spiral-disk galaxy. To
summarize the density-angular-momentum theory of galactic birth, if the protogalactic cloud happens to have high density and low
angular momentum, it will be born an elliptical galaxy, early on the Hubble sequence;
if the protogalactic cloud happens to have low
density and high angular momentum, it will be born a spiral-disk galaxy, late
on the Hubble sequence.
According to the
collision-merger theory of galactic birth, all galaxies are initially born spiral-disk
galaxies, forming from the collision and merger of several small protogalactic clouds.
Spiral-disk galaxies then grow larger by tearing apart and devouring
small satellite galaxies around it.
Finally, large elliptical galaxies result from the collision and merger
of two large spiral-disk galaxies. If
two spiral-disk galaxies happen to fall toward each other, they will eventually
collide. However, galaxies are not solid
objects; a galaxy is a collection of billions of star systems. Therefore, when two spiral-disk galaxies
collide, they actually pass through each other at first. However, the mutual gravity of the two
galaxies severely perturbs the orbits of the star systems in both galaxies,
causing the orbits of all the stars to become somewhat randomized and thus
disrupting the beautiful spiral arm structures of both galaxies. After the two galaxies pass through each
other, the two galaxies slow down, come to rest, and fall toward each other
again due to their mutual gravitational attraction. Again, the two galaxies pass through each
other, and again the mutual gravity of the two galaxies even more severely
perturbs the orbits of the stars in both galaxies, causing the orbits of all
the stars to become even more randomized and thus
further disrupting the beautiful spiral arm structures of both galaxies. Again, the two galaxies slow down, come to
rest, and fall toward each other yet again due to their mutual gravitational
attraction. Ultimately, the two galaxies
merge into a single galaxy with a round shape due to the randomized orbits of
all the stars. In other words, the two
spiral-disk galaxies have merged into an elliptical galaxy. Since galactic collisions result in galactic
mergers, astrophysicists regard a galactic collision to also
be a galactic merger. As we
discussed, our own Milky Way Galaxy and the Andromeda Galaxy are falling toward
each other, and they will collide in roughly five billion years. We now realize that these two spiral-disk
galaxies will not only collide, but they will also merge into an elliptical
galaxy. Astrophysicists have named this elliptical galaxy that will be born in roughly five
billion years the Milkdromeda Galaxy, since it will
be the merger of our Milky Way Galaxy and the Andromeda Galaxy.
There is strong evidence in
favor of the collision-merger theory of galactic birth. Firstly, the observable universe was smaller
billions of years ago, as we will discuss shortly. Therefore, galaxies must have been more
crowded together and hence collisions among them must have occurred more
frequently when they first formed billions of years ago. Secondly, our own Milky Way Galaxy is ripping
apart the small satellite galaxies around it as we discussed, and there is
evidence that some groups of stars within our Milky Way Galaxy were formerly
small satellite galaxies that our Milky Way Galaxy completely devoured as we
also discussed. Thirdly, computer
simulations reveal that the collision of two spiral-disk galaxies does indeed
result in a galactic merger into an elliptical galaxy. Finally, actual photographs of galactic
clusters reveal that the galaxies on the outskirts of galactic clusters are
predominantly spirals, while the galaxies toward the center of galactic
clusters are predominantly ellipticals. In fact, there is often a single giant
elliptical galaxy at the center of a galactic cluster. This distribution of spirals and ellipticals within galactic clusters suggests that all the
large galaxies in the galactic cluster were born spirals. Over billions of years, as spiral galaxies
fell toward the center of the galactic cluster, they collided and merged with
each other to form elliptical galaxies.
As these elliptical galaxies continued to fall toward the center of the
galactic cluster, they collided and merged with each other to form a giant
elliptical galaxy at the center of the galactic cluster. Although all of these observations and
calculations provide strong evidence in favor of the collision-merger theory of
galactic birth, there is also strong evidence against the collision-merger
theory of galactic birth. As we
discussed, spiral-disk galaxies have an abundance of interstellar gas and therefore have active star formation, while elliptical
galaxies have very little interstellar gas and therefore have very
little star formation, as we also discussed.
If two spiral-disk galaxies collide and merge, the interstellar gases
within these two spiral-disk galaxies should also collide. The collision of these gases should increase
their density and therefore should induce even greater star formation. Therefore, the collision-merger theory of
galactic birth predicts that two colliding spiral-disk galaxies with active
star formation should merge into an elliptical galaxy with even more active
star formation, but this is not correct.
Elliptical galaxies in fact have very little interstellar gas and thus
very little star formation, as we discussed.
This is a strong argument against the collision-merger theory of
galactic birth. We could modify this
collision-merger theory to claim that the collision of two spiral-disk galaxies
first results in a starburst galaxy, which is a large irregular galaxy with
much more active star formation than even spiral-disk galaxies. Presumably, most of the gas within a starburst
galaxy is consumed to form a large number of stars in a fairly
short period of time, leaving little gas to form further stars after
this relatively short duration of active star formation. Eventually, the randomized orbits of the
stars cause the entire galaxy to settle down into an elliptical galaxy, again
with a fairly round shape with little gas remaining to
form additional stars. Unfortunately,
this modification of the collision-merger theory to correct one false
prediction results in a new false prediction.
If elliptical galaxies were formerly starburst galaxies, then a large
number of stars of an elliptical galaxy should be young stars, but again this
is not correct. An elliptical galaxy is
an entire galaxy of mostly Population II stars, meaning that its stars are
mostly old stars with low metallicities.
This is another strong argument against the collision-merger theory of
galactic birth. The
density-angular-momentum theory of galactic birth seems reasonable, but it also
has strong counterarguments. According
to the density-angular-momentum theory of galactic birth, if the protogalactic cloud happened to have high density and low
angular momentum, it will be born an elliptical galaxy; if the protogalactic cloud happened to have low density and high
angular momentum, it will be born a spiral-disk galaxy. However, what happens if the protogalactic cloud happened to have high density and high
angular momentum? We could argue that
the galaxy would be born lenticular in this case. Lenticular galaxies have little gas and more
flat shapes. Presumably, the
high-density gas of the protogalactic cloud is consumed to form stars leaving little gas to form further
stars, and the high angular momentum of the protogalactic
cloud results in a more flat shape. This
is a strength of the density-angular-momentum theory over the collision-merger
theory, since the collision-merger theory offers conflicting explanations for
the formation of lenticular galaxies.
However, what prediction does the density-angular-momentum theory make
if the protogalactic cloud happened to have low
density and low angular momentum? Even
proponents of this density-angular-momentum theory of galactic birth do not
have a definitive answer to this question.
In summary, both of these theories of galactic birth each have their own
strengths and each have their own weaknesses.
Perhaps both theories are correct, since perhaps galactic birth occurs
through two different mechanisms.
Perhaps both of these theories are special cases of a more general
theory of galactic birth that has not yet been discovered. Perhaps both of these theories of galactic
birth are outright wrong, and perhaps a new theory of galactic birth that has not yet been discovered is the correct theory for the
formation of galaxies.
The most distant galaxies in
the observable universe are gigaparsecs (billions of
light-years) distant. These distances
are measured using redshifts and the Hubble law, as we will discuss
shortly. All of these extremely remote
galaxies have incredibly luminous centers.
Consequently, these distant galaxies are called
active galaxies, and their luminous centers are called active galactic nuclei,
which astrophysicists always abbreviate AGNs. A galaxy that does not have an active
galactic nucleus is classified as a quiet galaxy. There are no active galaxies within hundreds
of megaparsecs (several hundred million light-years)
from our galactic neighborhood; all galaxies within hundreds of megaparsecs (several hundred million light-years) of distance
from us are quiet galaxies. All galaxies
sufficiently distant from us are active galaxies with active galactic
nuclei. Although there are several
different types of active galaxies, the two most common types are Seyfert galaxies and quasi-stellar objects. Seyfert galaxies,
named for the American astronomer Carl Seyfert who
studied them, are active and distant galaxies as compared with quiet galaxies,
but Seyfert galaxies are not as active and not as
distant as quasi-stellar objects, which are the most active and the most
distant galaxies in the observable universe.
These quasi-stellar objects, often abbreviated QSOs and also referred to as simply quasars, are so distant
that they appear almost as point-like as stars even through powerful telescopes,
hence the name quasi-stellar. Active
galaxies are thousands of times more luminous than quiet galaxies, and much of
this luminosity is in the X-ray band of the Electromagnetic Spectrum. The active galactic nucleus of a quasar is so
luminous that it often outshines the entire host galaxy, preventing us from
even being able to observe the quasar’s host galaxy surrounding its active
galactic nucleus. Although Seyfert galaxies are very luminous, they are not as
luminous as quasars, and hence we are able to observe the host galaxy
surrounding the active galactic nucleus of a Seyfert
galaxy.
For a number of decades,
astrophysicists debated the source of the incredible energy that powers active
galactic nuclei. Astronomers made the
following observations of all active galactic nuclei. Firstly, there are variations in the
luminosity of active galactic nuclei over timescales shorter than one
year. Since the vacuum speed of light is
the speed limit of the universe according to relativity theory as we discussed
earlier in the course, variations in the luminosity of anything in the universe
places a constraint on the size of the luminous object. If active galactic nuclei have varying
luminosities on timescales shorter than one year, then the size of an active galactic
nucleus must be smaller than one light-year, perhaps the size of the Solar
System. In other words, an active
galactic nucleus is tiny compared to the overall size of its host galaxy. Secondly, we can calculate the mass of an
active galactic nucleus from the orbiting gases near the center of the host
galaxy. By measuring the blueshifts and the redshifts of the light from these
orbiting gases relative to the center of the active galactic nucleus, we can
determine their orbital speeds, the gravitational force responsible for these
orbital speeds, and hence the mass that exerts this gravitational force. We determine that the mass of a typical
active galactic nucleus is at least millions of solar masses. With this much mass crammed within such a
small region of space, astrophysicists were forced to conclude that a
supermassive black hole is the source of the incredible energy that powers
active galactic nuclei. Indeed, it is
not difficult to calculate that a black hole converts mass into energy with high
efficiency; just a few solar masses of material per year falling into a
supermassive black hole can provide sufficient energy to account for the
incredible luminosity of an active galactic nucleus. Thirdly, we often observe narrow columns or
jets of high-speed material from quasars.
These jets are often millions of parsecs (millions of light-years)
long! As the jets collide with the gases
that surround the active galaxy, radio waves are emitted,
resulting in enormous lobes of radio emissions surrounding the jets from some
quasars. As we discussed earlier in the
course, the source of the X-rays from X-ray binaries is an accretion disk
around a compact object. Moreover, some
of the gas that falls toward the compact object within an X-ray binary may be ejected as narrow columns or jets near the rotational
angular momentum axis of the accretion disk around the compact object. This makes some X-rays binaries similar to
quasars, but on a much smaller size scale than quasars. Indeed, some types of X-ray binaries are called microquasars. The presence of incredibly long jets from
quasars strongly suggests the presence of an enormous accretion disk around a
supermassive black hole. By combining
all of this observational and theoretical evidence, astrophysicists eventually
formulated the following model for active galactic nuclei: an enormous
accretion disk, perhaps the size of our Solar System, surrounding a
supermassive black hole. Caution: an
accretion disk the size of our Solar System is still tiny compared to the
overall size of the host galaxy. As the
gas within the accretion disk falls toward the supermassive black hole, the gas
is heated to millions of kelvins of temperature,
radiating an incredible amount of X-rays from the center of the galaxy, thus
powering the active galactic nucleus of the active galaxy.
Why are all
galaxies in the local universe quiet galaxies, and why are all extremely
remote galaxies active galaxies? What is
the difference between the local universe and the distant universe that makes
local galaxies and distant galaxies so different? Perhaps we are asking the wrong
question. We must realize that if we are
observing a galaxy ten billion light-years distant for example, this means that
it took its light ten billion years to travel from that galaxy to our Milky Way
Galaxy. This means that we are observing
that remote galaxy as it appeared ten billion years ago when it was extremely
young, presumably when it was first forming.
Hence, all galaxies sufficiently distant from us appear as they did when
they were first forming. Perhaps there
is no difference between the local universe and the distant universe. Perhaps all galaxies are born as active galaxies
with active galactic nuclei, and perhaps the supermassive black hole powering
the active galactic nucleus spends billions of years devouring the accretion
disk around it. As the supermassive
black hole devours the surrounding accretion disk, the active galactic nucleus
becomes more and more quiet, and perhaps the active galaxy gradually settles
down to become a quiet galaxy. The
distinction between Seyfert galaxies and quasars
supports this model of galactic evolution.
As we discussed, quasars are the most distant and the most active
galaxies, while Seyfert galaxies are less distant and
less active as compared with quasars.
Since quasars are the most distant galaxies, we are observing these
galaxies as they first formed. Since Seyfert galaxies are less distant, we are observing these
galaxies somewhat later in their evolution, after the supermassive black hole
has devoured a sufficient amount of the surrounding accretion disk that the
galaxy is somewhat less luminous as compared with its luminosity when the
galaxy first formed as a quasar. Further
evidence for this model of galactic evolution is the supermassive black hole at
the center of every major galaxy, including our own Milky Way Galaxy as we
discussed. Presumably, all quiet
galaxies, including our Milky Way Galaxy, were born with quasars. As the supermassive black hole devours the
surrounding accretion disk, the quasar becomes more and more quiet, becoming a Seyfert galaxy and then eventually settling down to become
a quiet galaxy. Recall that the center
of our Milky Way Galaxy is in the direction of the constellation Sagittarius
(the centaur archer). Astronomers have
discovered a radio source within the constellation Sagittarius that they have
named Sagittarius A, within which is a more distinct radio source astronomers
have named Sagittarius A* (pronounced Sagittarius A-star). This distinct radio source Sagittarius A*
surrounds the precise location of the supermassive black hole at our galactic
center. Radio waves are on the opposite
end of the Electromagnetic Spectrum from X-rays, as we discussed toward the beginning
of the course. Just as extremely hot
gases emit X-rays, extremely cool gases emit radio waves. The radio emissions from Sagittarius A* must
come from very cool gas falling toward the supermassive black hole at our
galactic center. This very cool gas must
be all that remains of the large, hot accretion disk that once powered the
active galactic nucleus when our Milky Way Galaxy was first born. All of this evidence has brought
astrophysicists to a consensus that all major galaxies are born as active
galaxies powered by an enormous and incredibly hot accretion disk around a
supermassive black hole. Over billions
of years, active galaxies become more and more quiet, transitioning from
quasars to Seyfert galaxies and eventually settling
down to become quiet galaxies. When we
observe an active galaxy billions of light-years distant, we must realize that
presently at this very moment, that galaxy is actually a quiet galaxy. If it is presently a quiet galaxy, then
intelligent life may exist on one of the planets orbiting one of the stars
within that galaxy. Perhaps those
intelligent lifeforms have even built telescopes, and if they point their
telescopes toward our Milky Way Galaxy, they would observe our Milky Way Galaxy
as an active galaxy! After all, if it
takes billions of years for light to travel from that galaxy to our Milky Way
Galaxy, then it also takes billions of years for light to travel from our Milky
Way Galaxy to that remote galaxy!
Therefore, those intelligent lifeforms would be observing our Milky Way
Galaxy as it first formed billions of years ago with a quasar! Right now at this very moment in the present
day, we observe that distant galaxy as a quasar and that distant galaxy
observes our Milky Way Galaxy as a quasar, even though both galaxies are
presently at this moment quiet galaxies!
Gamma-ray bursts, which
astronomers always abbreviate GRBs, are arguably the
single greatest mystery in all of astrophysics.
Hundreds of times every year, astronomers detect a burst of gamma-rays from outer space.
Astrophysicists formerly believed that these gamma-ray bursts come from
within our own Milky Way Galaxy. As we
discussed earlier in the course, astronomers detect sudden and intense X-ray
bursts from X-ray binaries within our Milky Way Galaxy, assuming the compact
object is a neutron star. Gamma-rays
have only a little bit more energy than X-rays, and so it would seem reasonable
to conclude that on occasion, X-ray binaries would also generate a sudden and
intense burst of gamma-rays. However,
the Compton gamma-ray observatory (CGRO) revealed
that gamma-ray bursts do not come from within our Milky Way Galaxy. If gamma-ray bursts came from within our
Milky Way Galaxy, then the distribution of gamma-ray bursts across the sky
would be concentrated along the band of milk that wraps around the entire sky,
the milky way.
However, when the Compton gamma-ray observatory mapped the distribution
of gamma-ray bursts across the sky, gamma-ray bursts were
revealed to come roughly equally from all directions in the sky. In other words, gamma-ray bursts are
extragalactic in origin. Gamma-ray
bursts come from extremely distant galaxies, from hundreds of millions to even
billions of light-years distant! Imagine
how powerful an explosion must be for us to still detect gamma-rays
at these incredible distances! The
supernova explosion of a high mass star is a weak explosion compared with these
incredible explosions!
If a particular gamma-ray
burst came from a galaxy billions of light-years distant, this means that it
took the gamma-rays billions of years to travel from
that galaxy to our Milky Way Galaxy.
After all, gamma-rays are a form of
electromagnetic radiation, a form of light that propagates at the vacuum speed
of light. Therefore, whatever explosion
caused the gamma-ray burst actually occurred billions of years ago, when the
universe was still very young. Often,
several hours after the gamma-ray burst, optical
telescopes detect an increase in visible light from the source galaxy of the
gamma-ray burst. This is
called the afterglow of the gamma-ray burst. The shape of the afterglow’s light curve is
nearly identical to the shape of the light curve of a Type II supernova
explosion. This strongly suggests that
the gamma-ray burst was caused by the death of a
high-mass star. Since the explosion that
caused the gamma-ray burst actually occurred billions of years ago when the
universe was still very young, we conclude that the source of the gamma-ray
burst and the subsequent afterglow was the death of an ancient Population III
star. As we discussed, Population III
stars were the first generation of stars born in the entire universe. As we also discussed, no Population III stars
have ever been discovered, suggesting that all
Population III stars were born high-mass main sequence stars with short
lifetimes. However, the explosion that
caused the gamma-ray burst is much more energetic than a supernova
explosion. We conclude that the first
generation of stars born in the universe, Population III stars with zero
metallicity, formed as not just high-mass stars, but as very high-mass stars,
perhaps with masses roughly equal to the Eddington limit, the theoretical
maximum mass of any star, as we discussed earlier in the course. After an incredibly short lifetime, perhaps
even shorter than one million years, these Eddington-limit stars swelled to
become larger and more luminous than even supergiant stars. These are called
hypergiant stars, which exploded with significantly greater luminosity than
even a supernova. This incredibly
violent explosion is called a hypernova. These hypernovae
are the most powerful explosions in the entire universe. We conclude that most
gamma-ray bursts are caused by the hypernova
explosions of ancient Population III hypergiant stars in extremely distant
galaxies. Several hypergiant
stars have been discovered in our Milky Way Galaxy,
but these are present-day Eddington-limit stars, not ancient Population III
Eddington-limit stars.
If all Population III stars
were born as very high-mass stars with masses roughly equal to the Eddington
limit, then we should be bombarded with many more
gamma-ray bursts from distant galaxies than we actually observe. We conclude that the hypernova
explosion of a hypergiant star causes a gamma-burst that does not radiate
spherically outward from the hypernova but is instead
concentrated into narrow beams from the hypernova. As such, exploding Population III hypergiant
stars in distant galaxies would emit gamma-rays bursts in particular
directions, and therefore most gamma-ray bursts would not be
ejected toward our general direction.
Hence, we only observe a small number of the gamma-ray bursts that
actually occurred in the ancient and young universe, the gamma-rays bursts that
happened to be ejected toward our general
direction. This would also explain the
extraordinary energy of gamma-ray bursts from such incredible distances. If a gamma-ray burst were
radiated spherically outward from a hypernova,
then its total energy would spread over a larger and larger sphere as it
propagates outward, diluting its energy as it travels billions of light-years
from the hypernova.
If instead the gamma-ray burst is concentrated
into narrow beams, then its total energy would not be significantly diluted as
it travels billions of light-years from the hypernova. Note however that the hypernova
also produces a more conventional Type II supernova explosion, emitting light
that does propagate spherically outward.
The light from this more conventional Type II supernova explosion
therefore does become diluted as it travels billions
of light-years, which we detect as the visible light afterglow a few hours
after the gamma-ray burst. In summary,
the first generation stars born in the entire universe, Population III stars
with zero metallicity, were born differently from later generations of stars and also died differently from later generations of
stars. Population III stars were all
born as very high mass stars with masses roughly equal to the Eddington
limit. That is, all Population III stars
were born hot, luminous, large, and high-mass with
short lifetimes. Population III stars
swelled to become hypergiant stars that died with hypernova
explosions, much more energetic than supernova explosions, producing gamma-ray
bursts concentrated along narrow beams followed by more typical Type II
supernova explosions that propagate spherically outward. The gamma-ray bursts and the associated
afterglows traveled across billions of light-years of space over billions of
years of time. Some of these gamma-ray
bursts happened to be ejected in our general
direction, which we observe when they finally arrive at our planet Earth in the
present-day universe.
Not all
gamma-ray bursts are caused by the hypernova
explosions of ancient Population III hypergiant stars. Some gamma-ray bursts are caused by the collision and merger of
binary neutron stars in distant galaxies. Since the spacetime curvature near a neutron star is nearly as
severe as the spacetime curvature near a black hole,
the collision and merger of binary neutron stars should create gravitational
waves that are similar to the collision and merger of binary black holes. However, electromagnetic waves (light) cannot
escape from within the event horizon of a black hole, as we discussed earlier
in the course. Therefore, the collision
and merger of binary black holes should not generate a gamma-ray burst. Electromagnetic waves (light) can however escape from neutron stars, and the collision and
merger of binary neutron stars should be violent enough to generate both
electromagnetic waves (light) and gravitational waves. Most of the gravitational waves that have been directly detected since the historic year 2015
have been from the collision and merger of binary black holes, as we discussed
earlier in the course. However, in the
year 2017, the first gravitational waves were directly
detected from the collision and merger of binary neutron stars. A few seconds after this gravitational wave
detection, a gamma-ray burst was detected, as we would
expect from the collision and merger of binary neutron stars. Moreover, a visible light afterglow was detected a couple days after the gamma-ray burst. Other gamma-ray bursts may
be caused by the collision and merger of a black-hole-neutron-star binary. In this case, the black hole rips apart and
devours the neutron star. The resulting
radiation would be similar to the resulting radiation from the collision and
merger of binary neutron stars: gravitational waves followed by a gamma-ray burst
followed by a visible light afterglow.
There are still other proposed mechanisms to explain other gamma-ray
bursts. In summary, there
are several different mechanisms that may cause gamma-ray bursts, and
astronomers must combine a variety of observational techniques to determine
which mechanism caused a particular gamma-ray burst.
Perhaps the most outrageous
mechanism that has been proposed to cause some
gamma-ray bursts is exploding microscopic primordial black holes. In the 1970s, the
physicist Jacob Bekenstein tried to formulate a
theory of the thermodynamics of black holes.
The physicist Stephen Hawking on the other hand claimed that there is no
theory of the thermodynamics of black holes.
Hawking argued that if there were such a theory, then black holes would
have a temperature that would cause them to radiate energy with a continuous
blackbody spectrum consistent with their temperature, but by definition
nothing can escape from a black hole. Of
course, we do detect X-rays from black holes, but these X-rays are emitted from gas falling toward the black hole before
crossing the event horizon. As long as
the gas has not yet passed the event horizon, we could still detect X-rays or
any other type of radiation from this gas that falls toward the black
hole. Once however the gas has crossed
the event horizon of the black hole, we can no longer detect any radiation,
since nothing can escape from within a black hole. Not even light can escape from within the
event horizon. This is why these objects
are named black holes, since they appear to be holes
that are black! Consider an isolated
black hole with no surrounding gases that could fall toward the black
hole. Such an isolated black hole should
not radiate any energy whatsoever.
Therefore, Hawking argued that an isolated black hole cannot
have a temperature; we could regard the temperature of an isolated black hole
as absolute zero temperature. Jacob Bekenstein disagreed with Stephen Hawking’s argument; Bekenstein instead argued that a theory of the
thermodynamics of black holes could actually be formulated. These two physicists made a friendly wager
between them. Bekenstein
wagered that there is a theory of the thermodynamics of black holes, while
Hawking wagered that there is no such theory.
As Stephen Hawking developed the theoretical physics necessary to win
this wager, he soon realized that Bekenstein was
correct; there is a theory of the thermodynamics of black holes. Stephen Hawking even succeeded in formulating
all of the mathematical details of black hole thermodynamics. At first, Hawking kept the results
private. Firstly, he did not want to
lose the wager! Secondly, he did not
believe his own results, at least at first.
Hawking calculated that an isolated black hole has a non-zero
temperature and therefore does indeed radiate energy with a continuous
blackbody spectrum consistent with its temperature, but how could this possibly
be the case? Eventually, Hawking
realized that his new theory is correct, but it must be
interpreted within the context of Relativistic Quantum Field Theory.
Several physicists were
involved in formulating Relativistic Quantum Field Theory, including American
physicist Richard Feynman, American physicist Julian Schwinger, Japanese
physicist Shin'ichirō Tomonaga,
and British-American physicist Freeman Dyson.
According to Relativistic Quantum Field Theory, there are particles that
continuously appear everywhere in the universe out of the vacuum of
nothingness. These are
called virtual particles. We
should all protest this theory, since virtual particles appearing everywhere
out of the vacuum of nothingness would violate fundamental laws of physics,
such as the conservation of mass and the conservation of energy. However, Relativistic Quantum Field Theory also
claims that these virtual particles disappear back into the vacuum of
nothingness before we are able to directly observe
their existence! Hence, the fundamental
laws of physics are not violated if we do not actually
observe any such violation! We are now
inclined to believe that Relativistic Quantum Field Theory is unscientific
nonsense, since it claims that fantastic things occur while ensuring that we
can never actually observe them occurring!
However, we can observe the effects of these virtual particles, even
though we cannot directly observe the virtual particles themselves. For example, two neutral metal slabs should
not attract or repel each other electromagnetically, since they are both
neutral. Of course, they do attract each
other gravitationally. However, virtual
particles that continuously appear and disappear around the metal slabs collide
with the metal slabs, exerting a pressure on both of them. Fewer virtual particles appear and disappear
between the two neutral metal slabs, since the boundary conditions imposed on
the partial differential equations yield solutions with quantized (discrete)
energies instead of a continuum of energies.
Therefore, the greater number of virtual particles that appear and
disappear on either side of the two neutral metal slabs exert a greater
pressure than the fewer number of virtual particles that appear and disappear
between the two neutral metal slabs.
Hence, the two neutral metal slabs feel a pressure that pushes them
toward each other! We can interpret this
pressure as an attraction in addition to their gravitational attraction. This is called the
Casimir effect, named for the Dutch physicist Hendrik Casimir who first
predicted this extra attraction from Relativistic Quantum Field Theory. This Casimir effect has
actually been observed in the laboratory. This is experimental evidence of the
existence of virtual particles, even though we cannot directly detect the
virtual particles themselves! Another
spectacular piece of evidence of the existence of these virtual particles
despite the fact that we cannot directly observe them is the anomalous g-factor of the electron. The g-factor
of any quantum-mechanical particle is the ratio between its quantum-mechanical
magnetic-moment-spin ratio and its classical magnetic-moment-spin ratio. Without these virtual particles, Relativistic
Quantum Mechanics predicts that the g-factor
of the electron should be exactly equal to two.
However, as virtual particles continuously appear and disappear around
an electron, they change its g-factor
slightly. If we experimentally measure
the g-factor of the electron, the
result is 2.002319304362; if we theoretically calculate the g-factor of the electron taking into
account the effects of virtual particles, the result it 2.002319304467, correct
to ten significant figures! No other
theory in the history of science is anywhere nearly this accurate. Among all theories among all the sciences,
nothing is as strongly experimentally verified as the
existence of these virtual particles that we cannot directly observe as
predicted by Relativistic Quantum Field Theory!
If virtual particles are
actually appearing and disappearing everywhere in the universe, Stephen Hawking
realized that they would be appearing and disappearing around the event horizon
of a black hole. Imagine virtual
particles appearing and disappearing in pairs around the event horizon of a
black hole. In some cases, both members
of the virtual pair disappear outside of the event horizon of the black
hole. In other cases, both members of
the virtual pair fall into the event horizon of the black hole. However, in some cases only one member of a
virtual pair may fall into the event horizon, leaving the other virtual
particle with no one to disappear with; hence, this virtual particle is forced to become a real particle. This would truly violate the fundamental laws
of physics, such as the conservation of mass and the conservation of
energy. The only way to preserve these
fundamental laws of physics is to claim that the virtual particle that fell
into the event horizon disappears with a small amount of the mass of the
singularity of the black hole. This
reduces the mass of the singularity by a small amount and shrinks the event
horizon by a small amount. Therefore, we
may interpret the virtual-converted-to-real particle outside of the black hole
as effectively coming from the black hole’s singularity, even though this is
not precisely what occurred. If virtual particles are indeed converting to
real particles everywhere outside the event horizon of a black hole while other
virtual particles fall into the black hole and disappear with a small amount of
the singularity’s mass, then there must be streams of real particles flying
away from the black hole as the black hole loses mass and therefore
energy! This compelled Stephen Hawking
to utter one of his most famous phrases, that even
isolated “black holes ain’t so black!” This stream of real particles radiating away
from an isolated black hole is called Hawking radiation. Although Hawking
radiation has never been observed, most physicists agree that Hawking’s theory
is correct. Hawking also calculated that
this Hawking radiation has a continuous blackbody spectrum; thus, there is a
temperature associated with this Hawking radiation. This temperature is called
the Hawking temperature. Hawking even
calculated the entropy of an isolated black hole. Entropy is another thermodynamic variable
related to temperature and energy. The
entropy of an isolated black hole is called the
Hawking entropy. Thus, Bekenstein won the wager when Hawking finally revealed his
theory of the thermodynamics of black holes.
This infuriated Bekenstein for the rest of his
life; although Bekenstein won the wager, he only won
the wager because of Hawking’s genius!
In brief, Hawking’s theory claims that an isolated black hole loses mass
and has a shrinking event horizon as it radiates Hawking radiation. Hence, these are called
evaporating black holes. Hawking also
calculated that the Hawking temperature of these evaporating black holes
actually becomes hotter and hotter as the isolated black hole loses more and
more mass. Thus, the Hawking radiation
becomes more and more luminous. At the
very end of their lives, these evaporating black holes should explode with
nearly infinite luminosity. These are called exploding black holes. If the luminosity of an exploding black hole is nearly infinite, we would detect gamma-ray bursts from
these exploding black holes even if they were billions of light-years
distant. Unfortunately, a stellar black
hole would take much longer than the current age of the universe to evaporate
and explode, and a supermassive black hole would take even longer to evaporate
and explode. However, a microscopic
black hole with an event horizon roughly the size of the nucleus of an atom
would only take roughly fourteen billion years to evaporate and explode; this
is roughly equal to the current age of the universe. The initial Hawking temperature of such a
microscopic black hole would be tremendously hot, but as we will discuss
shortly the universe was actually this hot shortly after the Big Bang. Therefore, these microscopic black holes
could have been born in the fires of the Big Bang. For this reason, microscopic black holes are also called primordial black holes. These microscopic primordial black holes
could have been born in the fires of the Big Bang, they could have spent the
past fourteen billion years evaporating, and they could be exploding right
now. We would then detect these
explosions as gamma-ray bursts.
Therefore, exploding microscopic primordial black holes could be the
source of some of the gamma-ray bursts that we continuously observe from
distant galaxies.
Cosmology and the History of the Universe
The word cosmos means
universe. Therefore, cosmology is the
study of the universe, and a cosmologist is someone who studies the
universe. Haven’t
we been studying cosmology throughout the entire course? Actually, throughout this entire course we
have been studying astronomy and astrophysics, which we strictly define as the
study of objects within the universe. In
other words, astronomy and astrophysics is the study of stars, planets, moons,
asteroids, comets, nebulae, star clusters, galaxies, galactic groups, galactic
clusters, and galactic superclusters. An
astronomer or an astrophysicist is someone who studies any of these objects
within the universe. Cosmology is the
study of the entire universe, the universe itself, and a cosmologist is someone
who studies the entire universe, the universe itself. Instead of studying the birth, life, and
death of planets, stars, and galaxies, a cosmologist studies the birth, life,
and death of the entire universe, the universe itself. Whenever we discuss properties of the entire
universe, such as its age, its size, or its temperature, we are studying
cosmology.
We begin our discussion of
cosmology with a seemingly simple question: why is the sky dark at night? The answer to this seemingly simple question
seems obvious at first: isn’t the sky dark at night
simply because our Sun is not in the sky?
As we reflect upon this question further, we realize that there is a
problem with this simplistic answer. If
the universe is infinitely large with infinitely many stars, then shouldn’t the light from all those infinitely many stars add
up to infinity thus making the nighttime sky infinitely bright? In fact, the daytime sky should also be
infinitely bright, shouldn’t it? Many students argue that the sky is dark at
night because most of the stars in the universe are so distant that they appear
very dim, but this argument is incorrect.
It does not matter how dim stars appear from their far distances;
infinitely many amounts of dim light should still add up to infinity. Some students argue that the gases that fill
outer space block the light from distant stars, but this argument is also
incorrect. Interstellar gases that
absorb an infinite amount of light will become hotter and hotter until an
equilibrium is established; the gases themselves begin to radiate as much light
as they absorb. Hence, we are back to
where we started: the nighttime sky (as well as the daytime sky) should be
infinitely bright. As we reflect upon
all of these arguments, we begin to truly wonder, why
is the sky dark at night? The first
person to ask this question in this meaningful way was the German astronomer
Heinrich Olbers.
Consequently, this is called the Olbers paradox: why is the sky dark at night? Heinrich Olbers
also discovered the asteroids Pallas and Vesta, which
we discussed earlier in the course. The Olbers paradox is resolved through Einstein’s General
Theory of Relativity. Recall that both
of Einstein’s theories of relativity (Special and General) reveal that there is
a speed limit of the universe, the vacuum speed of light. As we will discuss shortly, Einstein’s
General Theory of Relativity also reveals that the universe has a finite age;
in other words, the universe had a definite beginning at a finite time in the
past. A finite age of the universe
together with a finite speed limit of the universe together prevent anyone in
the universe from observing the entire universe. Let us make this argument more clear. The true age of the universe is roughly
fourteen billion years as we will discuss shortly, but let us suppose instead
that the age of the universe is fathomably younger, perhaps only one hundred
years old. In this case, we could not
see a star two hundred light-years distant for example, since it would take two
hundred years for light to travel from that star to us, but we are supposing
that the entire universe is only one hundred years old. In other words, there has not been sufficient
time in the entire history of the universe for light to travel from that star
to us, since we are supposing that the entire universe is only one hundred
years old. If the entire universe were
only one hundred years old, then the furthest stars we could see would be one
hundred light-years distant, since light would barely have sufficient time to
travel from those stars to us. The
furthest distance we would be able to observe would be one hundred light-years
in all directions away from us, and we would not be able to observe anything in
the entire universe further than one hundred light-years distant. We now realize that it would appear as if we
were at the center of a sphere that is one hundred light-years in radius, and
the finite age of the universe together with the finite speed limit of the
universe would forbid us from seeing anything in the entire universe beyond
that sphere. Any other observer living
on any other planet orbiting any other star within any other galaxy would
observe the same thing; they would appear to be at the center of a sphere that
is one hundred light-years in radius, and the universe forbids them from seeing
anything in the entire universe beyond their own sphere. We now realize that even if the universe is
infinitely large with infinitely many stars shining with light that adds to
infinite luminosity, the finite age of the universe together with the finite
speed limit of the universe together forbid us from observing the entire
universe. The laws of physics themselves
constrain us to only observe a finite part of the
universe within a spherical region centered on wherever we are located in the
universe. This finite part of the
universe that we are only permitted to observe is
called the observable universe. Wherever
we live in the entire universe, we appear to be at the center of our own
observable universe, and our own observable universe is always a sphere with a
radius in light-years equal to the age of the universe in years. The spherical edge of our own observable
universe is called the cosmic horizon, since it is
rather like the event horizon of a black hole.
Although, instead of being outside the event horizon of a black hole and
being forbidden from observing within that event horizon, we are inside the
cosmic horizon of our observable universe, and we are forbidden from observing
outside that horizon! Indeed, the
equations of General Relativity suggest that the entire universe is
mathematically identical to a black hole turned inside out! Also notice that
with every passing year, the universe is one year older. Hence, the cosmic horizon must grow by one
additional light-year in radius with each passing year. Thus, wherever we live in the universe, our
cosmic horizon must expand away from us at a speed of one light-year per
year. At what speed must we move to
cover a distance of one light-year in a time of one year? One light-year per year equals the vacuum
speed of light of course! We now
conclude that wherever we live in the universe, the cosmic horizon must expand
away from us at the speed of light. As
we will discuss shortly, the space within the cosmic horizon is also expanding
away from us at proportionally slower speeds.
The actual age of the universe is roughly fourteen billion years, as we
will discuss shortly. We conclude that
wherever we happen to be located within the entire universe, we appear to be at
the center of our own observable universe.
Wherever we happen to be located within the entire universe, the shape
of our own observable universe is a sphere roughly fourteen billion light-years
in radius with an edge (our own particular cosmic horizon) that expands away
from us at the vacuum speed of light.
The laws of physics constrain us to only observe
the stars and galaxies within our finite observable universe, even if the
entire universe is infinitely large with infinitely many stars shining with
light that adds to infinite luminosity.
The laws of physics forbid us from observing anything outside of our
cosmic horizon. When we add together all
of the light from all of the stars only within our observable universe, we
calculate a nighttime sky that is dark.
This is the resolution of the Olbers
paradox. In brief, the finite age of the
universe together with the finite speed limit of the universe together force
the observable universe to have a finite size with a finite number of stars
shining with a finite amount of light, even if the entire universe is
infinitely large with infinitely many stars shining with light that adds to
infinite luminosity.
The first person to tackle
these cosmological questions using advanced mathematics was Albert Einstein
using his General Theory of Relativity.
When Einstein solved the equations of his General Theory of Relativity
for the spacetime of the entire universe, the
equations reveal that our four-dimensional spacetime
actually had a beginning at a finite time in the past. Again, this resolves the Olbers
paradox. According to the equations of
General Relativity, the beginning of our four-dimensional spacetime
was a single event (a single mathematical point unified with a single instant
of time), and this single event that began our
four-dimensional spacetime had infinite spacetime curvature.
We could call this beginning the moment of creation, but all
cosmologists call this beginning the Big Bang.
It is a common misconception that the Big Bang was a violent explosion
within an empty universe. This is a
complete misunderstanding of the equations of General Relativity. To even imagine an
empty universe would be to presuppose that spacetime
already existed before the Big Bang, which is not correct. Again, according to the equations of General
Relativity, the Big Bang was the beginning of spacetime. Therefore, there was no spacetime
before the Big Bang. To
even imagine an empty universe before the Big Bang would be to imagine a
spacetime that existed before spacetime
began, which is obviously a contradiction!
There was absolutely nothing before the Big Bang, but by this nothing we do not mean an empty universe with nothing in
it. The four-dimensional spacetime itself was nonexistent! Many students demand an answer to the
following question: what happened before the Big Bang? This is a meaningless question! This question is just as meaningless as the
following question: what is north of the north terrestrial pole? There is nothing north of the north
terrestrial pole since that is the most northern point on planet Earth or any
other planet! There was nothing before the
Big Bang since the Big Bang was the beginning of spacetime
itself! It is meaningless to imagine
anything existing or occurring before the beginning of time itself. In other words, there was
no before that was before the Big Bang!
According to the equations of
General Relativity, spacetime itself expanded after
the Big Bang. We will discuss the origin
of matter and energy within spacetime shortly. For now, all matter within spacetime gravitationally attracts other matter within spacetime. This
mutual gravitational attraction should slow down the expansion of the
universe. Indeed, Einstein calculated that there are three different ways the universe could expand after
the Big Bang. If the density of
mass throughout the universe is greater than a certain
critical cosmic density, then the gravitational attraction among all this
matter would be strong enough to eventually overpower the expansion of the
universe. In this case, the universe
would eventually stop expanding and begin contracting until spacetime
ends at a single event (a single mathematical point unified with a single
instant of time), and this single event that would be
the end of spacetime has infinite spacetime
curvature. In other words, spacetime would end with an opposite of the Big Bang. All cosmologists call this end the Big
Crunch, since it is the opposite of the Big Bang. If the density of mass throughout the
universe is less than this certain critical cosmic density, then the
gravitational attraction among all matter within spacetime
would not be strong enough to eventually overpower the expansion of the universe. In this case, the universe would continue
expanding forever. If the density of
mass throughout the universe is exactly equal to this certain critical cosmic
density, the universe would also continue to expand forever but at a slower and slower rate due to the gravitational attraction
among all the mass within the universe.
According to the equations of General Relativity, these three possible
universes have different cosmic geometries.
If the density of mass throughout the universe is greater than the
critical cosmic density, then the cosmic geometry of the universe is
closed. In a closed geometry, the sum of
the angles in a triangle is greater than 180°, the circumference of a circle is
less than 2π multiplied by its radius, and lines that begin parallel do
not remain parallel but instead converge toward one another. A concrete example of a closed geometry is
the geometry of a sphere, which is itself a special case of the geometry of an
ellipsoid, which is also a closed geometry.
If the density of mass throughout the universe is less than this
critical cosmic density, then the cosmic geometry of the universe is open. In an open geometry, the sum of the angles in
a triangle is less than 180°, the circumference of a circle is greater than
2π multiplied by its radius, and lines that begin parallel do not remain
parallel but instead diverge away from one another. A concrete example of an open geometry is the
geometry of a hyperboloid. If the
density of mass throughout the universe is exactly equal to this critical
cosmic density, then the cosmic geometry of the universe is flat. A flat geometry is Euclidean (or normal)
geometry, where the sum of the angles in a triangle is equal to 180°, the
circumference of a circle is equal to 2π multiplied by its radius, and
lines that begin parallel do remain parallel to one another. A concrete example of a flat geometry is the
geometry of a plane. Students often
claim that our universe cannot have this third type of cosmic geometry. According to General Relativity, gravity is
the curvature of spacetime, as we discussed earlier
in the course. Therefore, many students
claim that a flat universe would have no curvature and therefore would have no
gravitation. However, the cosmic
geometry of the universe is the overall geometry of the entire universe. Even if the cosmic geometry of the universe
were flat, the gravitation of planets and stars and galaxies and black holes
within the universe create tiny curvatures within this overall flat
geometry. The same is true if the cosmic
geometry of the entire universe is closed or open. There are tiny gravitational curvatures
caused by planets and stars and galaxies and black holes, and these tiny
gravitational curvatures are superimposed upon the
cosmic geometry, which is the overall geometry of the entire universe.
Which of these three possible
universes do we live in? There is still
some debate among cosmologists on the answer to this question. If we add together the mass of all the normal
(luminous star) matter in the universe, the resulting density is much less than
the critical cosmic density. This would
suggest that we certainly live in an open universe that will continue to expand
forever. However, there is roughly ten
times as much dark matter as normal (luminous star) matter. This tremendous quantity of dark matter is
sufficient to cause us to live in a flat universe that will also continue to
expand forever. The dark matter could
possibly be sufficient to cause us to live in a closed universe that will not
continue to expand forever but will instead eventually begin contracting and
end with a Big Crunch. Again, it is
frustrating to have no idea what dark matter is composed of given its
extraordinary importance. Without dark
matter, stars within a galaxy would not remain bound within the galaxy, thus
causing individual galaxies to disperse.
Without dark matter, galaxies within a galactic cluster would not remain
bound within the galactic cluster, thus causing galactic clusters to
disperse. We now realize that without
dark matter, the entire universe would also disperse (expand forever)! The only size scale where dark matter is not
necessary is the star system scale. For
example, the dynamics of our Solar System is completely
explained through the gravity of the Sun, the planets, the moons, the
asteroids, the comets, and so on and so forth.
Without dark matter, our Solar System would not disperse; our Solar
System would remain together due to the gravitational attraction of our Sun.
In the early twentieth
century (the early 1900s), the scientific community
did not yet understand the crucial importance of the Big Bang, the beginning of
spacetime, in resolving the Olbers
paradox. Consequently, most physicists
during the early twentieth century (the early 1900s)
did not believe that the universe ever had a beginning. Einstein himself did not believe that the
universe ever had a beginning.
Consequently, he doubted all three cosmic solutions of his own General
Theory of Relativity, since all three solutions reveal that our
four-dimensional spacetime actually had a beginning
(the Big Bang) at a finite time in the past.
Einstein even tried to hide these three solutions by introducing a fudge
factor into his General Theory of Relativity, which he called the cosmological
constant. However, in the year 1929,
Edwin Hubble discovered that the universe is indeed expanding. Hubble discovered that the light from all
galaxies beyond our Local Galactic Group is redshifted, revealing that all
galaxies beyond our Local Galactic Group are moving away from us. Caution: galaxies within a galactic group or
within a galactic cluster actually fall toward each other due to their mutual
gravitational attraction. For example,
the light from the Andromeda Galaxy is actually blueshifted,
revealing that our Milky Way Galaxy and the Andromeda Galaxy are falling toward
each other, as we discussed. More
strictly, Hubble discovered that galactic groups and galactic clusters are
expanding away from each other, although even galactic groups and galactic
clusters may deviate from this cosmic expansion due to local gravitational
attractions. According to the equations
of General Relativity, this cosmic expansion is not an actual motion of galactic
groups and galactic clusters; the spacetime is itself
expanding and thus spacetime itself carries galactic
groups and galactic clusters away from each other. Nevertheless, this cosmic expansion manifests
itself as recessional motion, thus causing redshifted light as galactic groups
and galactic clusters recede from each other.
Because of this motion of all galaxies beyond our Local Galactic Group
away from us in all directions, Edwin Hubble announced to the world that the
universe is expanding. Hence, Edwin
Hubble was given credit for the discovery of the
expansion of the universe, even though this originally followed from Einstein’s
General Theory of Relativity. After
Hubble’s discovery, Einstein called his own doubts and his introduction of the
cosmological constant the “biggest blunder” of his life. Today, six persons are collectively given
credit for formulating the Big Bang model of cosmology, most importantly Edwin
Hubble for his observational work and Albert Einstein for discovering the
General Theory of Relativity and first mathematically deriving that our
four-dimensional spacetime had a beginning (the Big
Bang). The four cosmologists who further
developed the mathematical details of the expansion of the universe were the
Russian physicist Alexander Friedmann, the American
physicist Howard Robertson, the British mathematician Arthur Walker, and the
Belgian Catholic priest Georges Lemaître. The prediction of the expansion of the
universe is the first of the three great triumphs of the Big Bang model of
cosmology. We will discuss the other two
great triumphs of the Big Bang model of cosmology shortly.
Many students claim that if
all galaxies beyond our Local Galactic Group are expanding away from us,
doesn’t this prove that we are at the center of the universe? We must always keep in mind our earlier
discussion: wherever we happen to be located in the universe, it would appear
that we are the center of our own observable universe, and the universe appears
to expand away from our own particular location within the universe. This is the case with every observer in the
entire universe. There is no center of
the entire universe, since every point in the universe appears to be the center
of its own observable universe. As spacetime expands, every galactic group and galactic
cluster is carried away from every other galactic
group and galactic cluster. Therefore,
anyone living in any galactic group or galactic cluster in the entire universe
would observe all other galactic groups and galactic clusters expand away from
their own particular galactic group or galactic cluster. An analogy will help make this more clear. Imagine a
balloon inflated with enough air to give it a spherical shape, and imagine many
ants living on the outer surface of this spherical balloon. Suppose all of the ants decide to remain
stationary, meaning that the ants do not crawl.
Now suppose someone blows more air into the balloon. As the balloon is further inflated, the
material of the balloon stretches, thus causing all of the ants to be further
and further apart from one another.
Every one of these ants would see all the other ants apparently moving
away from them; therefore, each ant would believe itself to be the center of
the expansion. In actuality, none of the
ants is the center of the expansion since every ant observes itself to be the
center of the expansion. Notice also
that each ant would see all the other ants appearing to recede even though they
are not actually crawling. In actuality,
the material of the balloon is stretching, thus carrying all of the ants away
from each other. All of the ants are
analogous to galactic groups and galactic clusters, and the material of the
balloon is analogous to the spacetime itself. As spacetime
stretches, galactic groups and galactic clusters appear to move away from each
other. From within any galactic group or
galactic cluster anywhere in the entire universe, all other galactic groups and
galactic clusters appear to recede, causing their light to become redshifted. Imagine there is at least one civilization of
intelligent lifeforms living in every galactic group or galactic cluster in the
entire universe. Each one of these
civilizations would observe every other galactic group or galactic galactic
cluster appear to move away from them, as if their own particular galactic
group or galactic cluster was the center of the expansion of the entire
universe. In actuality, there is no
center of the expansion of the universe because every galactic group or
galactic cluster in the universe appears to be the center of its own observable
universe.
The Hubble law relates the
recessional speed of all galaxies beyond our Local Galactic Group to their
distance from us. If a galaxy beyond our
Local Galactic Group is a distance d
from us, then the speed v (for
velocity) with which the galaxy moves away from us is given by the Hubble law,
which states v = H0 d, where H0 is called the Hubble constant and is equal to
roughly seventy kilometers per second per megaparsec. In other words, the Hubble law states that
the speed with which galaxies beyond our Local Galactic Group move away from us
is directly proportional to their distance from us. We again emphasize that it
is the stretching of the spacetime that carries
galaxies away from us. Although
we do observe that the light from these distant galaxies is redshifted, these redshifts are caused by the stretching of the wavelength of
light as it journeys from distant galaxies toward us. The same would be observed
from every other galaxy in the universe.
Again, if we imagine that there is at least one civilization of
intelligent lifeforms living in every galactic group or galactic cluster in the
entire universe, each of these civilizations would observe redshifted light
from all galactic groups and galactic clusters outside of their own particular
galactic group or galactic cluster. These redshifts are caused by the stretching of the wavelength of
light as it journeys from one galactic group or galactic cluster to any other
galactic group or galactic cluster.
This is called cosmological redshift, the third type of redshift we have
discussed in this course, the other two being kinematic redshift and
gravitational redshift. Although this cosmological redshift is caused by the stretching of spacetime, we may nevertheless interpret this
redshift as a kinematic redshift, since it does appear that all galaxies beyond
our Local Galactic Group are moving away from us. Since we may interpret cosmological redshifts
as kinematic redshifts, we may calculate recessional speeds from these redshifts
that we measure for light from galaxies beyond our Local Galactic Group. The direct proportionality between this
recessional speed and distance according to the Hubble law is consistent with
the entire universe beginning with a Big Bang.
An analogy will help make this more clear. Imagine we are standing in an enormous
parking lot, and suppose we are surrounded by a circle of cars all driving directly away from us at sixty miles per hour,
and furthermore suppose that all of these cars are sixty miles distant from
us. We would conclude that all of these
cars began driving away from where we are standing one hour ago, since it takes
one hour for a car to drive a distance of sixty miles at a speed of sixty miles
per hour. Now suppose we are also surrounded
by an additional circle of cars all driving directly
away from us at 120 miles per hour, and furthermore suppose that all of these
cars are 120 miles distant from us. We
would again conclude that all of these cars began driving away from where we are
standing one hour ago, since it takes one hour for a car to drive a distance of
120 miles at a speed of 120 miles per hour.
Now suppose we are surrounded by yet another circle of cars all driving directly away from us at 180 miles per hour, and
furthermore suppose that all of these cars are 180 miles distant from us. We again conclude that all of these cars
began driving away from where we are standing one hour ago, since it takes one
hour for a car to drive a distance of 180 miles at a speed of 180 miles per
hour. As long as further and further
cars are driving faster and faster away from us in direct proportion to their
distance from us, then all cars at all distances began driving away from where
we are standing at the same time in the past.
According to the Hubble law v = H0 d, the recessional speeds of all galaxies (beyond our
Local Galactic Group) are directly proportional to their distances from
us. Therefore, all galaxies everywhere
in the entire universe began moving away from our Local Galactic Group at the
same moment in the past, the Big Bang or the beginning of spacetime
(the moment of creation). Warning: this
conclusion again tempts us to conclude that we are indeed at the center of the
entire universe. In actuality, the
stretching of spacetime carries every galactic group
or galactic cluster away from every other galactic group or galactic cluster,
causing every civilization of intelligent lifeforms across the entire universe
to deduce the same Hubble law. There is
no center of the expansion of the universe because every galactic group or
galactic cluster in the universe appears to be the center of its own observable
universe as the spacetime of the entire universe
stretches (expands). Since the Hubble
constant is roughly seventy kilometers per second per megaparsec,
every megaparsec of distance from us results in an
additional roughly seventy kilometers per second of speed away from us. In particular, a galaxy one megaparsec distant from us moves at roughly seventy
kilometers per second away from us, a galaxy two megaparsecs
distant from us moves at roughly 140 kilometers per second away from us, a
galaxy three megaparsecs distant from us moves at
roughly 210 kilometers per second away from us, and so on and so forth. Notice that further and further galaxies are
moving faster and faster away from us in direct proportion to their distance
from us, and therefore all galaxies everywhere in the entire universe began
moving away from us at the same moment in the past, the Big Bang or the beginning
of spacetime (the moment of creation).
Speed equals distance divided
by time; therefore, time equals distance divided by speed. If we solve the Hubble law v = H0 d for the Hubble constant H0, we deduce that the
Hubble constant equals speed divided by distance; that is, H0 = v / d. However, speed divided by distance is not
time, since time equals distance divided by speed. We conclude that the reciprocal of the Hubble
constant is time, since the reciprocal of the Hubble constant is indeed
distance divided by speed. We define the reciprocal of the Hubble constant to be the Hubble
time, and the Hubble time is a rough estimate of the age of the entire
universe, just as the ratio between the distance traveled by all the cars in
our imaginary parking lot to their speed is equal to the time all of them began
to drive away from where we are standing. The age of the entire universe is the amount
of time all galactic groups and galactic clusters have been expanding away from
each other since the Big Bang. Again,
the Hubble constant is roughly equal to seventy kilometers per second per megaparsec. If we
take the reciprocal of seventy kilometers per second per megaparsec
and perform a unit conversion, we calculate that the Hubble time is roughly
equal to fourteen billion years. We have
finally justified how we know the age of the entire universe. It truly is as simple as setting the speed of
galaxies equal to their distance traveled divided by the time that they have
been traveling! Note that the reciprocal
of a large number is a small number, and the reciprocal of a small number is a
large number. Thus, if the Hubble
constant is small, then the Hubble time is large; if the Hubble constant is
large, then the Hubble time is small.
This stands to reason. If the
universe is expanding slowly (small Hubble constant), then the universe must
have been expanding for a long duration of time since the Big Bang (large
Hubble time). If the universe is
expanding quickly (large Hubble constant), then the universe must have been
expanding for only a short duration of time since the Big Bang (small Hubble
time). Notice that as the universe ages,
the Hubble time must become larger and larger and therefore the Hubble constant
must become smaller and smaller.
Therefore, the Hubble constant is not truly a constant. Nevertheless, we would need to wait hundreds
of millions of years to notice a substantial change in its value. Therefore, it is appropriate to continue to
refer to H0
as the Hubble constant.
The Hubble law is the highest
rung of the Cosmological Distance Ladder.
To use this law to measure distances, first we determine the distances
to somewhat nearby galactic groups and galactic
clusters using lower rungs of the Cosmological Distance Ladder, such as the
Tully-Fisher relation, the Faber-Jackson relation, or the Type Ia supernova method.
We also determine the recessional speed of these galaxies by measuring
the redshift of their light. Caution: we
are interpreting a cosmological redshift as a kinematic redshift, as we
discussed. If we know the recessional
speed of galaxies and the distance to these galaxies, then the only unknown
remaining in the Hubble law v = H0 d is the Hubble constant H0. This is how we have determined that the
Hubble constant is roughly equal to seventy kilometers per second per megaparsec. To then
measure distances to incredibly remote galaxies, we simply measure the redshift
of their light. We interpret this
cosmological redshift as a kinematic redshift, and hence we calculate the recessional
speed of the galaxy. Since we have
already determined the Hubble constant H0, the only unknown remaining in the Hubble law v = H0 d is the distance.
This is how the Dutch astronomer Maarten Schmidt determined that quasars
are the most distant galaxies in the universe, by measuring their
redshifts. We now summarize the rungs of
the Cosmological Distance Ladder. The
lowest rung is the parallax method, which is only effective for nearby stars in
the solar neighborhood, out to distances of a couple thousand parsecs. The next rung is the main sequence fitting
method, which is effective for star clusters beyond the solar neighborhood but
still within our Milky Way Galaxy. The
next rung is the variable star method, which is effective for nearby galaxies
throughout the Local Galactic Group and even beyond the Local Galactic Group,
out to distances of a couple hundred megaparsecs. The next rung is the Tully-Fisher relation
and the Faber-Jackson relation, which is effective for even
more distant galaxies. The next
rung is the Type Ia supernova method, which is
effective for quite distant galaxies.
The highest rung of the Cosmological Distance Ladder is the Hubble law,
which is effective for the most distant galaxies in the observable universe.
The expansion of the universe
causes complications when discussing cosmological distances. As a concrete example, consider a galaxy that
is one billion light-years distant at a particular moment in time. This means that the light from that galaxy
would take one billion years to travel toward us, assuming the
one-billion-light-year distance remains fixed.
However, as the light traverses this distance, the universe expands,
stretching the distance that the light must traverse. Therefore, the light must actually travel
more than one billion light-years from a galaxy that was formerly one billion
light-years distant from us. Moreover,
the expansion of the universe also carries the distant galaxy away from us,
causing it to be even further from us than the distance traversed by the light
we receive from it. In summary, whenever
we observe a distant galaxy, its light has traversed a further distance than
the ancient distance that the galaxy was formerly from us when the light that
we presently observe first left the galaxy, and moreover
the galaxy is presently in actuality even further from us than even the
traversed distance of the light that we observe. For all of these reasons, cosmologists define
several different ways of measuring and calculating cosmological distances. In this course, we will simply define
cosmological distances in such a way that preserves the direct proportionality
between recessional speed and distance as determined by the Hubble law.
If the universe is expanding,
then it is gaining gravitational energy.
This increase in gravitational energy must come at the expense of
thermal energy, since energy must be conserved. Hence, the universe must become cooler and
cooler as it expands. This implies that
the early universe was hotter than the present universe. The observable universe expanded to its
present size from a formerly smaller size and cooled to its present temperature
from a formerly hotter temperature. Therefore,
the Big Bang model of cosmology is more properly called
the Hot Big Bang model of cosmology. We
can imagine traveling further and further backward in time when the observable
universe was smaller and smaller and thus hotter and hotter. When the universe was sufficiently young, the
observable universe may have been so small that all galaxies in the universe
were crowded against each other. At even
earlier times, galaxies did not even form yet; the entire universe was filled with stars that were relatively crowded
together. At even earlier times, not
even stars had formed; the entire universe was filled
with gas that would later form the first stars.
As we run the cosmic clock further and further backward in time, the gas
that filled the entire universe was hotter and hotter at earlier and earlier
times, when the universe was younger and younger. We can imagine times shortly after the Big
Bang when the observable universe was so incredibly small and so incredibly hot
that the laws of physics actually manifested themselves differently from our
present understanding of the laws of physics.
As we discussed earlier in the course, there are presently four
fundamental forces in our universe.
Listed in the correct order from strongest to weakest, these fundamental
forces are the strong nuclear force, the electromagnetic force, the weak
nuclear force, and the gravitational force.
Recall that the gravitational force is by far by far by far by far the
weakest force in the entire universe; the gravitational force is much much much
much weaker than the three other fundamental forces. In the 1970s,
the American physicist Sheldon Lee Glashow, the Pakistani physicist Abdus Salam, and the American physicist Steven Weinberg
formulated a theory claiming that at incredibly hot temperatures, the
electromagnetic force and the weak nuclear force unify into a single force that
they called the electroweak force, since it is the unification of the
electromagnetic force and the weak nuclear force. This theory is called
Relativistic Quantum Electroflavodynamics (or
Electroweak Theory for short). According
to this theory, at incredibly hot temperatures, there should only be three
fundamental forces in our universe: the strong nuclear force, the electroweak
force, and the gravitational force.
According to Relativistic Quantum Electroflavodynamics,
this electroweak unification occurs at a temperature of roughly three
quadrillion kelvins! Note that nowhere
in the entire present-day universe is it hot enough for this electroweak
unification to occur. The core
temperature of our Sun is roughly fifteen million kelvins; this is incredibly
hot by human standards, but this is also incredibly cold
compared with the electroweak unification temperature! The core of a helium-burning star is roughly
one hundred million kelvins, still too cold for electroweak unification! Temperatures at the center of a high mass
star, which then suffers a supernova explosion, are in the billions of kelvins,
still too cold for electroweak unification!
We might now suspect that Relativistic Quantum Electroflavodynamics
is a purely speculative theory that can never be tested
experimentally. However, electroweak
unification has been proven experimentally using subatomic particle
accelerators. A subatomic particle
accelerator uses electric and magnetic fields to accelerate subatomic particles
to incredibly fast speeds. In the 1980s, physicists succeeded in building subatomic particle
accelerators large enough with electric and magnetic fields strong enough to
accelerate protons and antiprotons to speeds unimaginably close to the speed
limit of the universe, the vacuum speed of light. These physicists then used these subatomic
particle accelerators to accelerate protons and antiprotons in opposite
directions and forced them to collide with each other at these incredible
speeds. These collisions were so
energetically violent that they effectively had a temperature of a few
quadrillion kelvins. As a result,
physicists actually witnessed, for just a fraction of an instant, electroweak
unification during these violently energetic proton-antiproton collisions! Glashow, Salam, and Weinberg received the
Nobel Prize in Physics for their correct theory. According to Relativistic Quantum Electroflavodynamics, the electroweak force divorces itself
into the electromagnetic force and the weak nuclear force below roughly three
quadrillion kelvins through a spontaneously broken symmetry involving the
Higgs-Englert particle, named for British physicist
Peter Higgs and Belgian physicist François Englert,
the two physicists who theorized the existence of this particle. Although electroweak unification was experimentally proven with subatomic particle
accelerators in the 1980s, these accelerators were
still not large enough to cause particle collisions energetically violent
enough to synthesize the Higgs-Englert particle. The largest particle accelerator in the world
is currently the Large Hadron Collider in Europe, abbreviated the LHC, and in the year 2012 this giant subatomic particle
accelerator caused particle collisions energetically violent enough to finally
synthesize the Higgs-Englert particle. Both Peter Higgs and François Englert received the Nobel Prize in Physics for their
correct prediction of the existence of this particle. Although the cores of stars and even
supernova explosions are too cold for electroweak unification to occur, humans
have built subatomic particle accelerators on planet Earth that have achieved
the electroweak unification temperature!
Although nowhere in the entire present-day universe is
it hot enough for electroweak unification to occur (other than subatomic
particle accelerators humans have built on planet Earth), there must have been
a very early time shortly after the Big Bang when the universe was so hot that
the entire universe only had three fundamental forces: the strong nuclear
force, the electroweak force, and the gravitational force.
Since Relativistic Quantum Electroflavodynamics has been experimentally proven, other
physicists have been inspired to search for theories that unify the strong
nuclear force with the electroweak force.
These theories are called grand unification
theories, which physicists abbreviate GUTs. According to these grand unification
theories, at an even hotter temperature far beyond the electroweak unification
temperature, a grand unification would result in only two fundamental forces in
the universe: a grand-unified force and the gravitational force. We will reveal the grand unification
temperature shortly. For now, we express
this grand unification threshold in terms of the appropriate size of a
subatomic particle accelerator to achieve grand unification. Again, the largest particle accelerator in
the world is currently the Large Hadron Collider in Europe, abbreviated the LHC. Unfortunately,
even the Large Hadron Collider is not large enough to achieve grand
unification. Although there are many
different grand unification theories that predict somewhat different grand
unification temperatures, all these different temperatures are roughly equal to
each other, and hence we can estimate the approximate size of a subatomic
particle accelerator that could achieve grand unification. According to most grand unification theories,
grand unification can only be achieved in a subatomic
particle accelerator roughly the size of our Milky Way Galaxy! Humans will never ever succeed in
constructing such an enormous subatomic particle accelerator. We might now suspect that grand unification
theories are purely speculative theories that can never be
tested experimentally. However,
there are other methods to test grand unification theories. According to these grand unification
theories, the proton is actually an unstable particle that disintegrates after
a certain lifetime. If we were to actually observe a proton disintegrate, this would be experimental
evidence that grand unification theories are correct. Unfortunately, we have never witnessed a
proton disintegrate. Perhaps grand
unification theories are correct that the proton is an unstable particle, but
perhaps the lifetime of a proton is much longer than the current age of the
universe; this would explain why we have never observed a proton
disintegrate. Nevertheless, if grand
unification theories are correct, then all the protons in the universe should
eventually disintegrate. Stars and planets and mountains and buildings and humans and
mobile telephones are all composed of atoms, which are in turn composed of
protons. If the lifetime of a proton is
perhaps fourteen billion years, the current age of the universe, then all the
protons in the universe could disintegrate at any moment. Perhaps tomorrow, all stars
and planets and mountains and buildings and humans and mobile telephones
will spontaneously disintegrate due to their unstable protons as predicted by
grand unification theories! The most
popular grand unification theory is Supersymmetric Relativistic Quantum Field Theory, or Supersymmetry for short. According to Supersymmetric Relativistic
Quantum Field Theory, for every particle of matter or antimatter in the
universe, there is a corresponding supersymmetric particle. For example, electrons, positrons
(antielectrons), neutrinos, and antineutrinos are all
classified as leptons, but Supersymmetric Relativistic Quantum Field
Theory claims that there are supersymmetric leptons called sleptons. This speculative theory also claims that
there are supersymmetric quarks called squarks,
supersymmetric photons called photinos,
supersymmetric gluons called gluinos, and
supersymmetric gravitons called gravitinos. Other supersymmetric particles include winos
and zinos. The
Large Hadron Collider might be large enough to create one of these
supersymmetric particles; this would be experimental evidence for a grand
unification theory. Unfortunately, we
have never discovered a supersymmetric particle. In other words, sleptons,
squarks, photinos, gluinos, gravitinos, winos, and zinos are all purely hypothetical particles. Without ever witnessing a proton disintegrate
and without ever discovering a supersymmetric particle, we currently have no
experimental evidence that any of the several grand unification theories are
correct. Nevertheless, most physicists
do believe in grand unification.
Although nowhere in the entire present-day
universe is it hot enough for grand unification to occur, there must have been
a very early time shortly after the Big Bang when the universe was so hot that
the entire universe only had two fundamental forces: the grand-unified force
and the gravitational force.
Although we have no
experimental evidence that any of the grand unification theories are correct,
some physicists claim that at an even hotter temperature the grand-unified
force unifies with the gravitational force.
Theories that claim that this occurs are called
super unification theories, which physicists abbreviate SUTs,
or theories of everything, which physicists abbreviate TOEs. A super unification theory or a theory of
everything would finally achieve Einstein’s ultimate dream to discover the
single ultimate equation that explains everything about the entire
universe. According to super unification
theories or theories of everything, at fantastically hot temperatures there
would be no gravity, no electromagnetism, and no nuclear forces. There would only be a single force throughout
the entire universe, called the super-unified force. We will reveal the super unification
temperature shortly. For now, we express
this super unification threshold in terms of the appropriate size of a
subatomic particle accelerator to achieve super unification. According to super unification theories,
super unification can only be achieved in a subatomic
particle accelerator roughly the size of our observable universe! Humans will never ever succeed in
constructing such an enormous subatomic particle accelerator. We might now suspect that super unification
theories are purely speculative theories that can never be
tested experimentally. However,
there are other methods to test super unification theories. Just as Relativistic Quantum Electrodynamics
reveals that light is composed of individual photons, super unification
theories or theories of everything claim that gravity is composed of individual
gravitons. If we could detect a single
graviton, this would be evidence for super unification theories or theories of
everything. Unfortunately, we have never
detected a single graviton.
Gravitational waves were just recently detected
for the first time in the year 2015, as we discussed earlier in the
course. The most popular super
unification theory or theory of everything is string theory, more properly
called brane theory. The word brane is a
shortening of the word membrane; a string is a one-dimensional membrane, a drum
is a two-dimensional membrane, and so on and so forth. This string theory or brane theory is also
called M-theory, meaning any of the following: membrane theory, matrix theory,
mystery theory, mysterious theory, magic theory, magical theory, magnificent
theory, majestic theory, momentous theory, monster theory, monstrous theory,
mother theory, or mother of all theories.
Some physicists jokingly suspect that the uppercase (capital) letter M
is actually an upside-down uppercase (capital) letter W for Witten, since the
American theoretical physicist Edward Witten has made the greatest mathematical
advances in this proposed super unification theory or theory of
everything. According to string theory
or brane theory or M-theory, every particle in the universe and even the
curvature of spacetime itself is actually composed of
vibrating branes (membranes) that are fantastically tiny, far far far far
smaller than the nucleus of an atom and even far far smaller than protons and
neutrons. The different vibrations of
these branes (membranes) would create different particles, such as leptons and
quarks and photons and gluons and even the speculative supersymmetric
particles. The graviton itself is a
vibrating brane (membrane) according to M-theory. We currently have no experimental evidence
that any of the several super unification theories are correct. Nevertheless, most physicists do believe in
super unification. Although nowhere in
the entire present-day universe is it hot enough for
super unification to occur, there must have been an incredibly early time
immediately after the Big Bang when the universe was so fantastically hot that
the entire universe only had one fundamental force: the super-unified force.
We can use the equations of
General Relativity to calculate the cooling temperature of the entire universe
as it expands. From these calculations,
cosmologists have divided the history of the entire universe into cosmic epochs
based on cosmic events that occurred throughout the entire universe. This is similar to human history. Historians have divided human history into
ages based on important events that occurred in human history, beginning with
the Stone Ages followed by the Bronze Age, the Iron Age, the Greco-Roman Ages,
the Middle Ages, and finally the Modern Ages. The history of any particular country or
culture is similarly divided into periods based on
important events. For example, American
history begins with the Pre-British Colonial Period followed by the British
Colonial Period, the Revolutionary Period, the Early Nineteenth Century Period,
the Sectional Crisis Period, the Late Nineteenth Century Period, the Early
Twentieth Century Period, the Superpower Period, and finally the Hyperpower Period.
Geologists have taken the entire history of planet Earth and divided it
into geologic eras, again based on important geological events. These geologic eras include the Paleozoic
Era, the Mesozoic Era, and the Cenozoic Era.
Similarly, cosmologists have divided the entire history of the universe
into cosmic epochs based on cosmic events that occurred throughout the entire
universe. We have arrived at the grand
finale of the entire course: a brief history of the entire universe.
The Big Bang occurred at time
zero, the beginning of spacetime. According to the equations of General
Relativity, the temperature and the energy density and
the spacetime curvature were all equal to infinity at
the Big Bang. Consequently, the
mathematical equations of General Relativity actually break down at the Big
Bang, meaning that the mathematical equations of General Relativity fail to
answer any questions about the moment of the Big Bang itself. What exactly occurred at the instant of the
Big Bang? Was it the hand of God? Science provides no answer.
The Planck Epoch occurred
from time zero to a time of roughly 5×10–44 seconds, roughly fifty
quadrillionths of one quadrillionth of one quadrillionth of one second after
God created the universe! During the
Planck Epoch, the temperature of the universe cooled from infinity down to
roughly 1032 kelvins, roughly one hundred nonillion kelvins! During this epoch, there was no gravity,
there was no electromagnetism, and there were no nuclear forces. There was only a single fundamental force in
the entire universe: the super-unified force.
Therefore, the universe must have been governed
by a super unification theory or theory of everything. This could have been M-theory (brane theory
or string theory), but we are not certain.
Thus, cosmologists are almost completely uncertain about what exactly
occurred during the Planck Epoch.
However, cosmologists do agree that at the end of the Planck Epoch, the
universe became so cool (roughly one hundred nonillion kelvins!) that the
super-unified force could no longer continue to exist. At the end of the Planck Epoch, the
super-unified force divorced itself into the grand-unified force and the
gravitational force. This was when
gravity was born, at the end of the Planck Epoch. Many cosmologists also agree that this
divorce liberated a tremendous number of gravitons that filled the entire
universe. This is
called the Cosmic Graviton Background Radiation. This Cosmic Graviton Background Radiation was
born at the end of the Planck Epoch at roughly one hundred nonillion kelvins of
temperature. However, fourteen billion
years of cosmic expansion has cooled this Cosmic Graviton Background Radiation
to nearly absolute zero temperature. We
will reveal the present-day temperature of this Cosmic Graviton Background
Radiation shortly.
The GUT Epoch occurred from a
time of roughly 5×10–44 seconds to a time of roughly 10–36
seconds, roughly one trillionth of one trillionth of one trillionth of one
second after God created the universe!
During the GUT Epoch, the temperature of the universe cooled from
roughly 1032 kelvins down to roughly 1029 kelvins,
roughly one hundred octillion kelvins!
During this epoch, there were two fundamental forces in the entire
universe: the grand-unified force and the gravitational force. Therefore, the universe must
have been governed by a grand unification theory to explain the
grand-unified force and Einstein’s General Relativity theory to explain the
gravitational force. The grand
unification theory that explained the grand-unified force could have been
Supersymmetric Relativistic Quantum Field Theory, but we are not certain. Thus, cosmologists are nearly completely
uncertain about what exactly occurred during the GUT Epoch. However, cosmologists do agree that at the
end of the GUT Epoch, the universe became so cool (roughly one hundred
octillion kelvins!) that the grand-unified force could no longer continue to
exist. At the end of the GUT Epoch, the
grand-unified force divorced itself into the strong nuclear force and the
electroweak force. Many cosmologists
also agree that a tremendous amount of energy was liberated
from this divorce, causing the universe to expand by a fantastic amount that it
would not have suffered otherwise. This is called the theory of inflation, and it is a recent
addition to the Hot Big Bang model of cosmology.
The Electroweak Epoch
occurred from a time of roughly 10–36 seconds to a time of roughly
10–11 seconds, roughly ten trillionths of one second after God
created the universe! During the
Electroweak Epoch, the temperature of the universe cooled from roughly 1029
kelvins down to roughly 3×1015 kelvins, roughly three quadrillion
kelvins! During this epoch, there were
three fundamental forces in the entire universe: the strong nuclear force, the
electroweak force, and the gravitational force.
This is the earliest cosmic epoch during which cosmologists are somewhat certain what precisely occurred. The gravitational force is explained by
General Relativity theory, formulated by Albert Einstein, the electroweak force
is explained by Relativistic Quantum Electroflavodynamics,
formulated by Sheldon Lee Glashow, Abdus Salam,
Steven Weinberg, Peter Higgs, and François Englert,
and the strong nuclear force is explained by Relativistic
Quantum Chromodynamics. Several
physicists were involved in formulating Relativistic Quantum Chromodynamics,
including American physicist Murray Gell-Mann, Russian-American physicist
George Zweig, German physicist Harald Fritzsch, Swiss
physicist Heinrich Leutwyler, and the three American
physicists David Gross, Frank Wilczek, and Hugh David
Politzer. At
the end of the Electroweak Epoch, the universe became so cool (roughly three
quadrillion kelvins!) that the electroweak force could no longer continue to
exist. At the end of the Electroweak
Epoch, the electroweak force divorced itself into the electromagnetic force and
the weak nuclear force. Thus, the four
fundamental forces that continue to exist throughout the universe today only
began to exist as four separate forces at the end of the Electroweak Epoch,
roughly ten trillionths of one second after the Big Bang, when the electroweak
force divorced itself into the electromagnetic force and the weak nuclear
force. The electroweak force would never
appear again until roughly fourteen billion years later on planet Earth when
humans built large subatomic particle accelerators!
The Particle Epoch occurred
from a time of roughly 10–11 seconds to a time of roughly 10–2
seconds, roughly one hundredth of one second after God created the
universe! During the Particle Epoch, the
temperature of the universe cooled from roughly 3×1015 kelvins down
to roughly 1011 kelvins, roughly one hundred billion kelvins! The Particle Epoch provided the appropriate
temperature to possibly create microscopic primordial
black holes that would spend the next fourteen billion years evaporating and
eventually exploding to possibly cause some gamma-ray bursts. The Particle Epoch certainly provided
appropriate temperatures for quarks and leptons and gluons to come into
existence out of the energy that filled the entire universe. The quarks and gluons then combined with one
another to form protons and neutrons.
Hence, the Particle Epoch is when normal matter came into
existence. Note that we have no idea
when dark matter came into existence, since we do not even know what composes
dark matter! If normal matter and normal
antimatter appeared in equal amounts during the Particle Epoch, they would have
completely annihilated each other, leaving no normal matter or normal
antimatter to eventually form stars and planets. Although physicists do not yet understand
why, we must nevertheless conclude that slightly more matter appeared than
antimatter during the Particle Epoch.
When normal matter annihilated with normal antimatter during the
Particle Epoch, there would have remained a tiny amount of leftover matter from
this cosmic annihilation. This tiny
amount of leftover matter would eventually form the stars and planets of all
the galaxies in the universe! Perhaps
there was in actuality slightly more antimatter than matter that appeared
during the Particle Epoch. After the
cosmic annihilation during the Particle Epoch, perhaps there instead remained a
tiny amount of leftover antimatter, and perhaps it was this tiny amount of
leftover antimatter that would eventually form the stars and planets of all the
galaxies in the universe. Perhaps our
Sun and our planet Earth and mountains and buildings and humans and mobile
telephones are actually composed of antimatter, and perhaps we have mistakenly
named these atoms as matter when in fact we are actually composed of
antimatter!
The Nucleosynthesis Epoch
occurred from a time of roughly one hundredth of one second to a time of
roughly three minutes after God created the universe. During the Nucleosynthesis Epoch, the
temperature of the universe cooled from roughly one hundred billion kelvins
down to roughly one billion kelvins.
During this Nucleosynthesis Epoch, protons and neutrons began to
combine, forming atomic nuclei. This is
why this is called the Nucleosynthesis Epoch, since
nuclei were synthesized during this cosmic epoch. Cosmological calculations indicate that
roughly three-quarters (roughly seventy-five percent) of the normal matter that
filled the universe were protons that remained alone,
separate from each other and separate from the neutrons. Cosmological calculations also indicate that
the remaining roughly one-quarter (roughly twenty-five percent) of the normal
matter that filled the universe were protons and neutrons that combined into
quadruplets, two protons and two neutrons fusing into a single nucleus. Recall that a single proton is the nucleus of
the hydrogen atom, and also recall that two protons
and two neutrons together form an alpha particle, the nucleus of the helium
atom. Therefore, these cosmological
calculations predict that the normal mass of the universe should be roughly
three-quarters (roughly seventy-five percent) hydrogen and roughly one-quarter
(roughly twenty-five percent) helium. As
we have discussed numerous times throughout the course, this is indeed the case. This is the second great triumph of the Hot
Big Bang model of cosmology, the explanation of the chemical composition of the
normal mass of the universe. Hence, the
universe became roughly three-quarters hydrogen and roughly one-quarter helium
between roughly one hundredth of one second and roughly three minutes after the
Big Bang. Caution: the universe was
still so hot that hydrogen and helium were not neutral atoms yet. The entire universe was
filled with a hot plasma of hydrogen nuclei (protons), helium nuclei
(alpha particles), electrons, and photons all colliding with each other. When protons and neutrons fused to form
helium nuclei, a tremendous number of neutrinos was liberated
that filled the entire universe. This is called the Cosmic Neutrino Background Radiation. This Cosmic Neutrino Background Radiation was
born during the Nucleosynthesis Epoch at billions of kelvins of
temperature. However, fourteen billion
years of cosmic expansion has cooled this Cosmic Neutrino Background Radiation
to nearly absolute zero temperature. We
will reveal the present-day temperature of this Cosmic Neutrino Background
Radiation shortly. Cosmological
calculations place constraints upon the relative abundances of the nuclei that
were synthesized during the Nucleosynthesis Epoch, and by combining these
calculations with our observations of the chemical composition of the normal
matter of the universe, we can estimate the amount of normal mass that fills
the entire universe. These calculations
and observations reveal that the normal matter that fills the universe should
have a mass of roughly one-tenth of the total mass that fills the
universe. Yet again, we are compelled to
conclude that the entire universe is composed of roughly ten times as much
mysterious dark matter as normal matter.
Unfortunately, this still does not reveal what actually composes dark
matter. Consequently, we have almost no
idea during which epoch of cosmic history the dark matter first formed.
The Epoch of Nuclei occurred
from a time of roughly three minutes to a time of roughly three hundred
thousand years after the Big Bang.
During the Epoch of Nuclei, the temperature of the universe cooled from
roughly one billion kelvins down to roughly 3240 kelvins. During this Epoch of Nuclei, the entire
universe was filled with a hot plasma of colliding
nuclei, electrons, and photons. This hot
plasma cooled as the universe expanded.
In addition, the entire universe was filled
with the Cosmic Graviton Background Radiation and the Cosmic Neutrino
Background Radiation. Both of these
background radiations continued to cool as the universe expanded.
The Recombination Epoch
occurred from a time of roughly three hundred thousand years to a time of
roughly four hundred thousand years after the Big Bang. During the Recombination Epoch, the
temperature of the universe cooled from roughly 3240 kelvins down to roughly
2710 kelvins. We will regard roughly
three thousand kelvins as the average temperature of the universe during the
Recombination Epoch. During this cosmic
epoch, electrons combined with nuclei to form neutral hydrogen and helium
atoms. This is why this epoch is called the Recombination Epoch, since electrons combined
with nuclei during this cosmic epoch.
When the electrons combined with the nuclei to form neutral atoms, a tremendous
number of photons was liberated that filled the entire
universe. This is
called the Cosmic Photon Background Radiation. This Cosmic Photon Background Radiation was
born during the Recombination Epoch at roughly three thousand kelvins of
temperature. However, fourteen billion
years of cosmic expansion has cooled this Cosmic Photon Background Radiation to
nearly absolute zero temperature.
Cosmological calculations reveal that the present-day temperature of the
Cosmic Photon Background Radiation should be a miserable three kelvins above
absolute zero. It is easy to use the
Wien displacement law to calculate that at such an incredibly cold temperature,
the Cosmic Photon Background Radiation should have a continuous blackbody
spectrum with its primary radiation within the microwave band of the
Electromagnetic Spectrum. In the year
1964, the American astronomers Arno Allan Penzias and Robert Woodrow Wilson
built a microwave telescope in New Jersey.
They became frustrated however, since their microwave telescope
continuously detected microwaves coming from all directions in the sky with a
temperature of roughly three kelvins above absolute zero. It was only later that other astronomers and
cosmologists realized that Penzias and Wilson had accidentally discovered the
Cosmic Photon Background Radiation that fills the entire universe. Penzias and Wilson received the Nobel Prize
in Physics for this tremendous, although accidental, achievement. This is the third great triumph of the Hot
Big Bang model of cosmology, the prediction of the three-kelvin Cosmic Photon
Background Radiation that fills the entire universe. The Cosmic Background Explorer was NASA’s
great microwave space telescope, in operation from 1989 to 1993. This microwave telescope mapped the Cosmic
Photon Background Radiation to fair resolution.
The Cosmic Background Explorer was replaced by the
Wilkinson telescope, in operation from 2001 to 2010, which mapped the Cosmic
Photon Background Radiation to incredible resolution. We may interpret this map as an actual image
of how the entire universe appeared during the Recombination Epoch, between
roughly three hundred thousand years and roughly four hundred thousand years
after the Big Bang.
Although the Cosmic Photon
Background Radiation is nearly perfectly uniform, the
map constructed by the Wilkinson telescope reveals microkelvin
variations throughout the universe during the Recombination Epoch. Variations in temperature must correspond
with variations in density. Some regions
of the early universe were more dense than average, while other regions of the
early universe were less dense than average.
The regions of the universe that were more dense
than average must have collapsed under their self-gravity, ultimately forming
cosmic filaments. This would leave more
empty space between cosmic filaments, ultimately becoming cosmic supervoids. However,
computer simulations reveal that more dense regions would have had insufficient
self-gravity to collapse and form cosmic filaments without roughly ten times as
much mass as the normal matter that fills the universe. Once again, we conclude that roughly ninety
percent of the mass of the universe is the mysterious dark matter. Therefore, the variations in density that
existed during the Recombination Epoch were due primarily to variations in the
density of dark matter. Whatever
composes this mysterious dark matter, we are forced to
conclude that it already existed before the Recombination Epoch, since the microkelvin variations mapped by the Wilkinson telescope
reveal that dark-matter-density variations already existed during the
Recombination Epoch. Over billions of
years, galactic superclusters formed within cosmic filaments, galactic groups
and galactic clusters formed within galactic superclusters, galaxies formed
within galactic groups and galactic clusters, and stars formed within
galaxies. The first generation of stars
born were Population III stars with zero metallicity, being composed of pure
hydrogen and helium. These stars fused
some of their hydrogen to form more helium.
These Population III stars were very high mass stars that swelled to
become hypergiant stars that suffered hypernova-supernova
explosions, synthesizing all the atoms on the entire Periodic Table of Elements
and ejecting hot, rapidly-expanding supernova remnants that polluted or
enriched the surrounding universe with these metals. These gases eventually formed
second-generation Population II stars with small but non-zero metallicity. These stars fused some of their hydrogen to
form more helium. Some of these stars
were high mass stars that ended their lives with supernova explosions,
synthesizing all the atoms on the entire Periodic Table of Elements and
ejecting hot, rapidly-expanding supernova remnants that further polluted or
enriched the surrounding universe with even more of these metals. These gases eventually formed
third-generation Population I stars with higher metallicities than Population
II stars. The formation of nuclei by
stars is called stellar nucleosynthesis, which
continues to occur to the present day.
The formation of nuclei during the Nucleosynthesis Epoch is called Big Bang nucleosynthesis or primordial
nucleosynthesis, which only occurred from a time of roughly one hundredth of
one second to a time of roughly three minutes after the Big Bang. Billions of years of stellar nucleosynthesis
has increased the fraction (percentage) of helium and decreased the fraction
(percentage) of hydrogen throughout the universe. Nevertheless, fourteen billion years of
stellar nucleosynthesis has only changed these fractions (percentages) by small
amounts. The normal (atomic) mass of the
universe remains roughly three-quarters (roughly seventy-five percent)
hydrogen, roughly one-quarter (roughly twenty-five percent) helium, and a tiny
fraction (tiny percentage) of metals.
Over billions of years of cosmic history, the universe continued to
expand, causing galactic groups, galactic clusters, galactic superclusters, and
cosmic filaments to move away from each other and also
causing the three cosmic background radiations (photon, neutrino, and graviton)
to continue to cool.
Presently, the universe is
roughly fourteen billion years old. The
universe is filled with a Cosmic Photon Background
Radiation at a miserable three kelvins above absolute zero. The universe is also filled
with a Cosmic Neutrino Background Radiation, and cosmological calculations
reveal that this radiation should be at a miserable two kelvins above absolute
zero. This Cosmic Neutrino Background
Radiation has not yet been detected. If it is someday detected
and if it is measured to have a temperature of roughly two kelvins above
absolute zero, this will become the fourth great triumph of the Hot Big Bang
model of cosmology. Again, this has not yet been achieved.
If it is achieved, astronomers will use neutrino telescopes to construct
a map of this Cosmic Neutrino Background Radiation, providing a neutrino image
of how the entire universe appeared during the Nucleosynthesis Epoch, between
roughly one hundredth of one second and roughly three minutes after the Big
Bang, since that was when the Cosmic Neutrino Background Radiation was
born. The universe is
also filled with a Cosmic Graviton Background Radiation, and
cosmological calculations reveal that this radiation should be at a miserable
one kelvin above absolute zero. This
Cosmic Graviton Background Radiation has not yet been
detected. If it is someday detected and if it is measured to have a
temperature of roughly one kelvin above absolute zero, this will become the
fifth great triumph of the Hot Big Bang model of cosmology. Again, this has not yet
been achieved. If
it is achieved, astronomers will use graviton telescopes to construct a map of
this Cosmic Graviton Background Radiation, providing a graviton image of how
the entire universe appeared at the end of the Planck Epoch, roughly fifty
quadrillionths of one quadrillionth of one quadrillionth of one second after
the Big Bang, since that was when the Cosmic Graviton Background Radiation was
born. We may never achieve this
fifth triumph of the Hot Big Bang model of cosmology. As we discussed, we have never detected a
single graviton, and gravitational waves were just recently
detected for the first time in the year 2015.
We now summarize the entire
universe, and in doing so we will also summarize the entire course. The universe is filled
with a Cosmic Photon Background Radiation at a miserable three kelvins above
absolute zero, but this background radiation was born during the Recombination
Epoch at roughly three thousand kelvins of temperature when the universe was
between roughly three hundred thousand years old and roughly four hundred
thousand years old. The universe is
filled with a Cosmic Neutrino Background Radiation at a miserable two kelvins
above absolute zero, but this background radiation was born during the
Nucleosynthesis Epoch at billions of kelvins of temperature when the universe
was between roughly one hundredth of one second old and roughly three minutes
old. The universe is
filled with a Cosmic Graviton Background Radiation at a miserable one
kelvin above absolute zero, but this background radiation was born at the end
of the Planck Epoch at roughly one hundred nonillion kelvins of temperature
when the universe was roughly fifty quadrillionths of one quadrillionth of one
quadrillionth of one second old. Roughly
ninety percent of the mass of the universe is dark matter, which exerts normal
gravity even though it is not composed of normal (atomic) matter. The remaining roughly ten percent of the mass
of the universe is normal (atomic) matter.
Roughly three-quarters (roughly seventy-five percent) of this normal
(atomic) matter is hydrogen, roughly one-quarter (roughly twenty-five percent)
of this normal (atomic) matter is helium, and a tiny fraction (tiny percentage)
of this normal (atomic) matter is metals.
Most of this normal (atomic) matter is intergalactic/intracluster
gases within galactic clusters and interstellar/intragalactic
gases within galaxies, and some of these gases have formed stars that fuse
hydrogen into helium, increasing the amount of helium and decreasing the amount
of hydrogen in the universe by small amounts.
High mass stars synthesize metals.
A tiny fraction (tiny percentage) of this normal (atomic) matter has
formed planets, moons, asteroids, and comets.
Stars are clumped together to form galaxies, galaxies are clumped
together to form galactic groups and galactic clusters, galactic groups and
galactic clusters are clumped together to form galactic superclusters, and
galactic superclusters are clumped together to form cosmic filaments. There are roughly one hundred billion star
systems within a typical galaxy, and there are roughly one hundred billion galaxies
in the observable universe. Therefore,
there are roughly ten sextillion star systems in the observable universe. The entire universe has been expanding for
roughly fourteen billion years, carrying all galactic groups, galactic
clusters, galactic superclusters, and cosmic filaments away from each other and also causing all three cosmic background radiations
(photon, neutrino, and graviton) to cool.
The universe presently has four fundamental forces, but it was born with
only one fundamental force described by a super unification theory or a theory
of everything.
There is
one final topic about the universe that we must discuss. This final
topic is fascinating, arguably mysterious, and perhaps even beautiful. Somewhere within this photon and neutrino and
graviton filled, dark matter and hydrogen and helium dominated, continuously
expanding and cooling universe, there is a seemingly ordinary cosmic
filament. Within that seemingly ordinary
cosmic filament, there is a seemingly ordinary galactic supercluster. Within that seemingly ordinary galactic
supercluster, there is a seemingly ordinary galactic group. Within that seemingly ordinary galactic
group, there is a seemingly ordinary disk galaxy. Within that seemingly ordinary disk galaxy,
there is a seemingly ordinary spiral arm.
Within that seemingly ordinary spiral arm, there is a seemingly ordinary
middle-aged main sequence star, but this is not an ordinary middle-aged main
sequence star at all. This middle-aged
main sequence star is extraordinary, because the third planet orbiting that
star actually has life upon it. If this weren’t unbelievable enough, on a seemingly ordinary
continent on that extraordinary planet, there is a seemingly ordinary school,
with a seemingly ordinary building, with a seemingly ordinary classroom, but
this is not an ordinary classroom at all.
This is an extraordinary classroom on that extraordinary planet orbiting
that extraordinary star, because there is actually a person at the front of
that classroom who has explained the nature and the history of the entire
universe to a group of students, and perhaps this is the most fantastic thing
about the entire universe.
Libarid A. Maljian homepage at the Department of Physics at CSLA at NJIT
Libarid A. Maljian profile at the Department of Physics at CSLA at NJIT
Department of Physics at CSLA at NJIT
College of Science and Liberal Arts at NJIT
New Jersey Institute of Technology
This webpage was most recently modified on Wednesday, the fourth day of December, anno Domini MMXXIV, at 04:15 ante meridiem EST.