You are on page 1of 108

1/3/2016

Introduction to Physical Astronomy - The Sun

The Sun
The home page for the Solar Data Analysis Center at NASA Goddard Space Flight Center includes current solar images in a variety of wavelengths. Since
wavelength is inversely proportional to temperature, visible light images show the Sun's photosphere ("surface"), ultraviolet images show the chromosphere and the
transition region (20,000 to 1,000,000 K) to the corona, which is seen in x-ray images (and during eclipses):
The solar corona has very low density (less than a billion particles per cubic centimeter), but very high temperatures (up to 4 million K). (source)

(View Cosmos DVD 5, episode 9, on solar activity)


Here are a number of movies of the Sun. Look for granules (convection cells with dark edges where cooler gases sink), spicules (vertical "spikes"), coronal holes
(large areas of low density where the Sun's magnetic field opens outward, allowing outflow of the solar wind) and filaments ("curves" following magnetic field lines
above the surface):
a movie of the solar surface (10.59 MB) (source);
a Hinode movie of the same (15.86 MB) (source);
a STEREO movie of solar prominence on 4/13/10 (8.57 MB) (source);
an SDO (Solar Dynamics Observatory) closeup of an eruption on 12/31/12 (1.08 MB) (source);
an SDO movie of coronal rain following a coronal mass ejection on 7/19/12 (15.38 MB, centered on 304 Angstroms, covering 21.5 hours; source);
a SOHO movie of sunspots from the week starting 3/27/01 (1.06 MB) (source);
a SOHO movie of solar activity, in UV, from 10/21/03 to 11/06/03 (24.61 MB) (source);
Hinode magnetic field images (4.46 MB) from a 12/13/06 flare (2.81 MB) (source);
a Hinode movie of solar X-ray jets (26.14 MB) (source).
a STEREO movie of a coronal mass ejection washing over Earth (40.52 MB) (source).
The magnetic fields of astronomical bodies are thought to be generated by the movement of conducting fluids. In the case of the Earth, this is the liquid Iron in the
outer core, and results in the reasonably regular dipole field we observe, with its North and South magnetic poles.
The case of the Sun, however, is much more complicated. In addition to the Sun's differential rotation, in which the rotation speeds vary with latitude, the conducting
plasma responsible for the Sun's magnetic field undergoes radial as well as polar convection in the outer 30% of the Sun's radius. The result is much more chaotic,
where field lines are warped and often reconnect with violent results:
A snapshot of the solar magnetic field. (source)

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20The%20Sun.html

1/5

1/3/2016

Introduction to Physical Astronomy - The Sun

Especially strong magnetic fields inhibit convection, and so their poles on the solar surface are cooler and appear darker (although the regions surrounding them are
usually much brighter in ultraviolet). These are the familiar sunspots.
Sunspots come and go in 11 year cycles:
The most recent complete solar cycle (seen here in 284 Angstroms) began in 1996. (source)

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20The%20Sun.html

2/5

1/3/2016

Introduction to Physical Astronomy - The Sun

Note the Maunder Minimum between 1645 and 1715:


Sunspot cycles since 1610. (source)

At the start of each new cycle, the polarity of sunspot "pairs" reverse. For instance, during the cycle which is beginning in 2008, active regions in the
northern hemisphere are oriented so that their south poles are west of their north poles (the opposite occurs in the southern hemisphere). In the previous
cycle, the order was reversed. Here is a SOHO movie of magnetic fields from May 1998 (8.37 MB) (source).
The solar wind, particles streaming from the surface of the Sun, gives us auroras and comet tails. It is a ubiquitous phenomenon, especially in hot young stars like LL
Orionis:
Bow shock around LL Orionis. (source)

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20The%20Sun.html

3/5

1/3/2016

Introduction to Physical Astronomy - The Sun

The solar wind speed is around 400 km/s, which is about 2/3 of the escape velocity at the surface of the Sun. The wind, however, starts from the
corona, which extends several solar radii; the escape velocity at 2.4 solar radii is 400 km/s. The current solar wind velocity is available from SOHO.
The solar wind carries away about 10-14 solar masses annually. This is about 634 million kg per second. If the wind was entirely made up of protons,
the outgoing flux would be 3.79 * 1035 protons per second. Assuming that they are emitted isotropically, we should detect an average of about 3.37 per
cubic centimeter at the Earth. The measured value is actually about twice that.
Even radiation pressure sculpts the interstellar medium:
Cloud IC 349, shaped by radiation pressure from the star Merope, above upper right. (source)

Solar flares are classified using the following scheme of letters and numbers (n):
class
An
Bn

Peak Intensity
n*

A-class flares are approximately at background levels.

n*

10-7 W/m2

B- and
C-class flares are not strong enough to have much of an effect on the Earth.
M-class flares can cause polar radio blackouts, and can pose a health risk to astronauts.

n*

10-6 W/m2

Mn

n*

10-5 W/m2

Xn

n * 10-4 W/m2

Cn

comments

10-8 W/m2

X-class flares cause intense auroras and can damage electromagnetic equipment.

Intensities are measured at wavelengths between 1 and 8 Angstroms.


Table of Contents
References
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20The%20Sun.html

4/5

1/3/2016

Introduction to Physical Astronomy - The Sun

Index
2013, Kenneth R. Koehler. All Rights Reserved. This document may be freely reproduced provided that this copyright notice is included.
Please send comments or suggestions to the author.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20The%20Sun.html

5/5

1/3/2016

Introduction to Physical Astronomy - Distance vs. Direction

Distance vs. Direction


Because the universe is so vast, many of the distance units we use are based on the travel time of the fastest thing there is: light. The speed of light in vacuum is
c = 299,792,458 meters per second
= 2.998 * 108 m/s.
(Scientific notation, using exponents, are ubiquitous in physics; the scope of our inquiry demands their use.)
The distance in meters can then be computed using
distance = velocity * time.
So, for example, one lightsecond is 2.998 * 108 m. We will find the following distance units useful:
One Astronomical Unit (AU) is the average radius of Earth's orbit, equal to 1.496 * 1011 meters (m).
One lightyear (ly) is the distance light travels in a vacuum in one year, equal to 9.461 * 1015 m.
One parsec (pc) is the distance to a star whose parallax angle is one arcsecond; 1 pc = 3.086 * 1016 m = 3.262 ly.
Here are some images which might help you appreciate the orders of magnitude involved in our study:
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows system, download the Java VM (Virtual Machine) for your
version of Windows at the download section at java.sun.com.
"NGC" stands for New Galactic Catalog, and objects designed by "M" followed by a number from 1 to 110 are Messier Objects, cataloged by Charles Messier
in the latter part of the 18th century.
(View Cosmos DVD 6, episode 10, Milky Way Simulation and Large Scale Structure Images.)

Parallax
The parallax of the red star is larger than the parallax of the gold star.

The two blue-green circles represent the Earth at opposing points in its orbit. The large yellow circle between them of course represents Sol (our Sun). Suppose we
want to measure the distances to the red and gold stars. As the Earth moves over half its orbit, each star appears to change position. This is called parallax and the
change in position is twice the parallax angle. The parallax angle for the red star is and that for the gold star is . is less than because the gold star is farther
away.
If we know the radius r of the Earth's orbit we can compute the distances Ri to the stars:
Rred = r cot () and Rgold = r cot ()
Since for large distances cot() is very close to 1/ (if is measured in radians!), we have
Rred = r / and Rgold = r /

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Distance%20vs.%20Direction.html

1/5

1/3/2016

Introduction to Physical Astronomy - Distance vs. Direction

A parallax angle of 1 arcsecond yields a distance of 1 parsec (there are 360 degrees around a circle, 60 minutes of arc in a degree and 60 seconds of arc in a
minute). Therefore to find the distance in parsecs of a star whose parallax is measured in milliarcseconds (mas), simply divide 1000 by the parallax.

The Partial Perspective Viewer


will help you find your place in the universe (with a very fond wink at the shade of Douglas Adams):
This applet knows where some things in the universe are, and will display their locations in 3D. You may change the orientation by dragging on the image: horizontal
dragging rotates about the z axis, verticle dragging tilts the z axis. You may change the distance scales using the "Zoom In" and "Zoom Out" buttons below. The Milky
Way Galaxy is identified at its center. Earth (representing the entire solar system) is drawn in green. Even though there is no "center" of the universe, Earth is drawn
for convenience at the origin of the graph.
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows system, download the Java VM (Virtual Machine) for your
version of Windows at the download section at java.sun.com.
As you zoom out into the furthest reaches of the universe, you will notice a "dust" of unlabeled dots. Each dot is a cluster of galaxies, from the Abell
Catalog. Some Abell clusters have been labeled, but most are not for the sake of clarity.
2MASS, the 2 Micron All Sky Survey, permits us to see the nearest million and a half galaxies all at once. (source)

Portfolio Exercise: If we adopt a scale of 1 cm to represent 1 lightsecond, then 1 foot would represent 30.5 lightseconds and 1 mile would represent
160934 lightseconds, or 1.86 lightdays.
If we imagine the Sun to lie at mile marker zero on I-75 in Florida, how far from that point would the Earth lie? The midpoint between Jupiter and
Saturn? Pluto?
I-75 is 467 miles long in Florida; 353 miles long in Georgia; 160 miles long in Tennessee; 192 miles long in Kentucky; 210 miles long in Ohio; and 394
miles long in Michigan. Where along I-75 would you locate Alpha Centauri?
To deal with the rest of the universe, we obviously need a different scale. Let us now use 1 inch to represent 1 parsec. Then 1 mile would represent
63360 parsecs. Where along I-75 (using this new scale) is Betelgeuse? The Crab Nebula? The center of the Milky Way? Supernova 1987A? M31?
M104? Stephen's Quintet? Place these seven objects as accurately as possible on this map of I-75 and include it in your portfolio.

The Changing Cosmos


We rarely see change as we observe the distant cosmos, supernovae being the obvious exceptions.
This time lapse of M3 over a single night shows one of the few phenomena that are observable on human time scales: variable stars of the RR Lyrae type. (source)

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Distance%20vs.%20Direction.html

2/5

1/3/2016

Introduction to Physical Astronomy - Distance vs. Direction

RR Lyrae is a star in the constellation Lyra. It is a variable star, whose luminosity changes regularly with time as it fuses Helium in its core. Its distance is known
because it belongs to a cluster whose parallax is known. Since stars of its type have a well-defined relationship between their luminosity and the period of their
variability, they are useful standard candles: objects whose behavior dictates their luminosity. This, as we will see, can be compared to their apparent brightness to
determine their distance.
This sequence of the Cat's Eye Nebula shows the expansion of the nebula over a 3 year period. (source)

Orion - What you see and where things really are


The constellation Orion; the red line represents the celestial equator, and the white line represents 6 hours Right Ascension.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Distance%20vs.%20Direction.html

3/5

1/3/2016

Introduction to Physical Astronomy - Distance vs. Direction

The stars are labeled with Greek letters (in order, "alpha", "beta", "gamma", "delta", "epsilon", "kappa" and "zeta") indicating their relative brightness
within the constellation; their common names, celestial coordinates, visual magnitude, parallax, distance and spectral classes are given here:
Name

Common Name

RA

dec

mag

Parallax (mas)

distance (ly)

Spectral Class

Orionis

Betelgeuse

05 55 10.29

+07 24 25.3

0.45 7.63

427.3

M2Ib

Orionis

Rigel

05 14 32.27

-08 12 05.9

0.18 4.22

772.5

B8Ia

Orionis

Bellatrix

05 25 07.87

+06 20 59.0

1.64 13.42

242.9

B2III

Orionis

Mintaka

05 32 00.40

-00 17 56.7

2.25 3.56

915.7

O9.5II+

Orionis

Alnilam

05 36 12.81

-01 12 06.9

1.69 2.43

1341.6

B0Ia

Orionis

Saiph

05 47 45.39

-09 40 10.6

2.07 4.52

721.2

B0.5Iavar

Orionis

Alnitak

05 40 45.52

-01 56 33.3

1.74 3.99

817.0

O9.5Ib SB

This data comes from the Hipparcos Catalog. Note that the common names are all of Arabic origin. Alpha Orionis is a semi-regular pulsating star,
gamma and kappa Orionis are variable stars, delta Orionis is an eclipsing binary and zeta Orionis is a double star.
This photograph of the Eastern horizon was taken in the early hours of a mid-October morning in 2002. It is a composite of several 8 second exposures taken with a
Canon G1 (in Cincinnati, it takes a little over 10 seconds for the Earth to rotate enough to cause a star to start to become a streak).
Note the tilt of the celestial equator, which is the projection of the Earth's equatorial plane into the sky: 23.45 degrees because of the tilt of the Earth's axis of
rotation. The declination (dec) measures "celestial latitude" relative to the celestial equator and is measured in degrees. Comparing the labeled image with the table
below, we can see that positive declinations are above the equator, and negative declinations are below it.
Also note that the Right Ascension (RA) increases from right to left; this is because angles increase in counterclockwise rotation, which when looking toward the
South is from West to East. The Right Ascension is measured in hours (24 hours is a circle, so there are 15 degrees per hour).
We will discuss magnitude and spectral class later, but there is one very important thing to notice about the table: the distances in lightyears. The constellation Orion
looks like it does to us only in a small neighborhood of the Earth. Since the stars are at widely varying distances, any significant shift in position will change the angles
between them, and the constellation will appear much different. You can use the Partial Perspective Viewer to get an idea of how this works: try zooming out to 102.5
parsecs and tilting the z axis. From that perspective, Rigel, Mintaka, Saiph and Alnitak form a trapezoid obviously disconnected from the others.
Here is a deeper view of Orion (and check out this image of Barnard's Loop).
Portfolio Exercise: Use SIMBAD to find the celestial coordinates of the five brightest stars in the constellation Cassiopeia ( through Cas). On a
piece of graph paper, set up a coordinate system where the horizontal axis runs from 0 to 2 hours of Right Ascension, and the vertical axis runs from 55
to 65 degrees of declination. Plot the locations of the five stars. Find the constellation in the night sky and use your graph to identify the stars. How do
you have to orient your graph so that it looks correct? Why did you have to orient it that way?

Astronomical Distances
Light transit time and parallax are just two of a number of methods for measuring astronomical distances; we will discuss several in more depth later:
technique
radar ranging

equation
d = c * echo time / 2

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Distance%20vs.%20Direction.html

approximate maximum
effective distance
Solar System
4/5

1/3/2016

Introduction to Physical Astronomy - Distance vs. Direction

stellar parallax

dpc = 1000 /
parallaxmas

H-R Diagram (main sequence spectral class determines absolute magnitude)

dpc = 10(mag - Mag +

RR Lyrae / Cepheid Variables: Period/Luminosity relation, calibrated to


known distance to Cepheids in the Large Magellanic Cloud

Mag ~ 1.608 *
perioddays 0.275

5)/5

200 pc

5000 pc
20 Mpc (~ Virgo Cluster)

Tully-Fisher Relation: absolute magnitude of a spiral galaxy (a and b depend Mag = a - b Log
100 Mpc
on morphological type and wavelength)
(width of spectral line)
Type Ia Supernovae: C-O White Dwarfs explode with characteristic
luminosity as a function of their light curves

dpc = 10(mag - Mag +

Hubble Relation between distance and recessional velocity

d = v / H0

5)/5

1000 Mpc
z= 1

There are numerous variations on these themes, some involving very clever uses of statistics which are beyond the scope of this text.
All of the techniques based on luminosity measurement depend on our knowledge of the extinction: how interstellar and intergalactic dust and gas absorb light.
Note that these techniques must be calibrated to each other: radar ranging is necessary to accurately determine 1 AU, which is needed for stellar parallax,
which is used to calibrate the magnitude-distance relation, etc.
Each technique has limits to its accuracy, for some approaching 20%, and they should in principle all agree where they overlap. At present that agreement is
better than 10%. That it is not zero is a testament to the difficulty of consistently measuring many of these quantities.
The maximum distance for the Hubble Relation is listed as "z=1". This refers to the red shift, about which more later.
Table of Contents
References
Index
2010, Kenneth R. Koehler. All Rights Reserved. This document may be freely reproduced provided that this copyright notice is included.
Please send comments or suggestions to the author.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Distance%20vs.%20Direction.html

5/5

1/3/2016

Introduction to Physical Astronomy - Astronomical Observation

Astronomical Observation
As we have seen, we can only directly measure the angle between two objects, not their distances from us. But if we know
their distance from us, we can compute the distance between the two objects. More usually, we are measuring the size of a
single object from the angle it subtends:
size = distance * angular diameterradians
where a radian is 180/ degrees. Here are some examples of variation of angular diameter with distance:
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows system, download
the Java VM (Virtual Machine) for your version of Windows at the download section at java.sun.com.
Perigee is the point of closest approach during the orbit of a satellite about the Earth. Similarly, perihelion is
the point of closest approach during orbit of the Sun. Apogee is the point of greatest distance from the Earth
during orbit, and aphelion is the point of greatest distance from the Sun during orbit.
By setting the angular diameter in the equation above to the resolution of a telescope, we can compute the size of the
smallest object it can see at any given distance. The resolution is determined by interference effects and depends on the size
of the optics (usually a mirror) and the wavelength being observed:
angular resolutionarcsec = .0025 * wavelengthAngstroms / mirror diametercm
For instance, the Hubble Space Telescope (below) has a mirror diameter of 2.4 meters, or 240 cm. At optical wavelengths
(for instance, 5600 Angstroms), its angular resolution would be
.0025 * 5600 / 240 = .0583 arc seconds,
times / (180 * 3600) radians in an arc second = 2.828 * 10-7 radians.
Looking at Saturn at closest approach (8.004 AU, or 1.197 * 1012 m), the Hubble could resolve an object whose
size = 1.197 * 1012 * 2.828 * 10-7 = 338.6 km wide.
The light gathering power (literally, how much electromagnetic radiation the instrument can gather in a given amount of time)
is proportional to the square of the mirror radius. This is important because intensity decreases with the square of the
distance: if you move twice as far away, the intensity drops to one quarter of what it was at the original position. This is why
we distinguish between apparent magnitude and absolute magnitude (M):
M = apparent magnitude + 5 - 5 * log10 distancepc.
Absolute magnitude is the magnitude a star would have at a fixed distance of 10 parsecs; the apparent magnitude (m) will
be larger than M if the star is more than 10 parsecs distant, and less than M if it is closer. You can barely see a star of
apparent magnitude 6 with your eyes, if you are in a region of dark skies. Under the same conditions, binoculars takes you
to magnitude 10 and 8 inch telescopes to about magnitude 13. The Hubble telescope can just see objects of apparent
magnitude 30.
There are severe limits to which types of electromagnetic radiation can be observed from the Earth's surface. The Earth's
atmosphere is transparent to radio waves with wavelengths between about 10 meters and 1 cm, and to infrared and visible
wavelengths from 105 Angstroms to the near ultraviolet (around 2900 Angstroms).
Turbulence in the atmosphere blurs point-like Betelgeuse into a "seeing disc" (1/10 speed) (source):

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Astronomical%20Observation.html

1/7

1/3/2016

Introduction to Physical Astronomy - Astronomical Observation

The degree of the effect is referred to as "good" or "poor" seeing. For these reasons, many modern observatories are
based in space.
Much of what we observe requires extremely long exposure times.
This image of the Orion-Eridanus Superbubble required 40 hours of exposure time. (source)

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Astronomical%20Observation.html

2/7

1/3/2016

Introduction to Physical Astronomy - Astronomical Observation

Telescopes
One of the most common telescope designs in use today is the Ritchey-Chretien. (source)

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Astronomical%20Observation.html

3/7

1/3/2016

Introduction to Physical Astronomy - Astronomical Observation

The Ritchey-Chretien design is a variation on the basic reflector telescope, and differs from the Cassegrain reflector in
that the main mirror is shaped like a hyperbola instead of like a parabola, improving the image quality.
Some of the premier telescopes in recent use are:
The Hubble Space Telescope orbits the Earth every 96 minutes at an altitude of 575 km. It is 13.2 m long and has a
mass of 11,110 kg. It is a Ritchey-Chretien design with a 2.4 m mirror and resolution of .05 as. It provides data at
the rate of approximately 120 GB/week. Hubble includes the following instruments:
The Advanced Camera for Surveys, has one 16 megapixel CCD (3500 to 11000 A) and two 1 megapixel
CCDs (1700 to 11000 A and 1150 to 1700 A), covering fields of view of 202x202, 26x29 and 31x35
arcseconds, respectively.
The Wide Field and Planetary Camera (3), has 48 filters, a 16 megapixel UV/visible CCD (2000 to 10000 A)
and a 1 megapixel near-IR detector (8000 to 17000 A), covering a field of view of 160x160 arcseconds.
The Cosmic Origins Spectrograph, sensitive to 1150 to 3000 A.
The Space Telescope Imaging Spectrograph, observing in the range 1150-10000 A with three 1 megapixel
detectors.
The Near Infrared Camera and Multi-Object Spectrometer, observing in the range 8000-25000 A.
The Fine Guidance Sensors, sensitive to 4000-7000 A, are not primarily for observation but have been used
for precision location measurement.
Galex: the Galaxy Evolution Explorer, orbits the Earth every 90 minutes at an altitude of 690 km. It is 2 m long and
has a mass of 280 kg. It is a Ritchey-Chretien design with a .5 m mirror and resolution of 5 as. It provides data at the
rate of approximately 1 TB/month.
It has two instruments: NUV, observing in the range 1750-2800 A, and FUV, observing in the range 1350-1750 A;
both have 2 Mpixel detectors.
The Chandra X-Ray Observatory, orbits the Earth every 64 hrs. 18 min. in an eccentric orbit ranging from 10000140161 km. It is 13.8 m long and has a mass of 4800 kg. It features a High Resolution Mirror Assembly with .84 m
optics and resolution of .5 as. It provides data at the rate of approximately 2.25 GB/week. Chandra includes the
following instruments:
The Advanced CCD Imaging Spectrometer, observing in the range .2-10 keV (1-62 A).
The High Resolution Camera.
The High Energy Transmission Grating, observing in the range .4-10 keV (1-30 A).
The Low Energy Transmission Grating, observing in the range 5-140 A.
The Spitzer Space Telescope trails Earth around Sun. It is 4m long and has a mass of 865 kg. It is a RitcheyChretien with .85 m optics and resolution of 1.5 as. Spitzer includes the following instruments:
The Infrared Array Camera, observing at 3.6, 4.5, 5.8 and 8 microns (1 micron = 104 A). Each detector has
65536 pixels, with resolution of 1.2 as/pixel.
The Infrared Spectrograph, observing at 5.2-14 (lo-res), 9.9-19.5 (hi-res), 14-38 (lo-res) and 19-37.2 (hires) microns.
The Multiband Imaging Photometer, observing at 24 (2.5 as/pixel resolution), 70 (9.8 or 4.9 as/pixel), 160
microns (16 as/pixel), and measuring 52-97 micron spectra.
The SOHO observatory has 12 instruments which continuously monitor the Sun, the solar corona and the solar wind.
The twin STEREO observatories each have four instruments which continuously monitor the Sun and its corona.
The Green Bank Telescope is a radio telescope with a diameter of 100 m and a mass of 7,300,000 kg. It features
16 receivers covering frequencies from 342 MHz to 115 GHz.
The VLA: Very Large Array includes 27 25 m dishes yielding a maximum resolution equivalent to a 36 km antenna,
with the sensitivity of a 130 m antenna. Each dish weighs 230 tons. The VLA has a resolution of .04 as at 43 GHz,
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Astronomical%20Observation.html

4/7

1/3/2016

Introduction to Physical Astronomy - Astronomical Observation

and features 8 receivers covering frequencies from 73 MHz to 50 GHz (at wavelengths of 400, 90, 20, 6, 3.6, 2, 1.3
and .7 cm).
The VLBA: Very Long Baseline Array includes 10 25 m dishes, each weighing 240 tons. It features 10 receivers
covering frequencies from 312 MHz to 90 GHz (the central frequencies are .326, .611, 1.438/1.658, 2.275, 4.999,
8.425, 15.369, 22.236, 43.174 and 86.2 GHz; the wavelengths are 90, 50, 21/18, 13, 6, 4, 2, 1, .7, .3 cm, and the
resolutions are 22, 12, 5/4.3, 3.2, 1.4, 0.85, 0.47, 0.32, .17, .1 mas).
Here are sample images from these telescopes:
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows system, download
the Java VM (Virtual Machine) for your version of Windows at the download section at java.sun.com.
And here is a movie of M 20 (1.74 Mb), taken by Spitzer's Multiband Imaging Photometer. (source).
(View Cosmos DVD 6, episode 10, dramatization of red shift discoveries at Mt. Wilson.)
Portfolio Exercise: Research 3 additional telescopes, each in a different region of the electromagnetic
spectrum.
Note that many telescopes, both ground-based and space-based, have multiple instruments. Be
specific about which instruments you are discussing.
For each, identify the range(s) of wavelengths it is sensitive to, and its resolution at those wavelengths. Use
published resolutions if you can find them; otherwise, compute the resolution from the equation above. What is
the smallest object it could discern at 1 parsec? 1000 parsecs? 1 million parsecs? For 1 and 1000 parsecs,
compute your answers in both parsecs and A.U.
For each telescope, include an image in representative colors taken by the telescope, and explain what
observational wavelengths correspond to each primary color.
Include references for wavelength and resolution information, and for the images.

Amateur Observing
So what can you expect to see through a reasonably good amateur telescope? Unfortunately, not as much as you might
expect. Here are sample images from an 8 inch Meade LX-90, using a 12.4 mm eyepiece:
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows system, download
the Java VM (Virtual Machine) for your version of Windows at the download section at java.sun.com.
These images were taken in suburban Cincinnati on a clear night, through the eyepiece with a Cannon G-1 and an Orion
SteadyPix camera mount. When taking photographs through the lens, alignment is a major headache. However, these
photos give a pretty good idea of what you see with your eye. While detail is lacking and the colors are somewhat washed
out, there is certainly something very exciting about seeing light from the planets with your own eyes, and no one fails to
express awe when they look at Saturn. But what about deep sky objects?
Looking through this telescope, M-31 (Andromeda) is a faint smear of dust. M-42 (the Orion Nebula) has a clear shape
but no color, and the LCD screen on the camera is not sensitive enough to focus properly at any rate. So why can't you see
what the Hubble sees? Even though you are on the ground, you'd think you could do better than that!
The whole problem is one of intensity. Your eye can't add up the light received over a long period, as a digital camera
might. When the intensity is so low, your color-sensitive cone cells are inactive, and so you only see in black and white. The
beautiful deep-sky photos shown by the Hubble were long exposures, with the light accumulated over minutes or hours; for
instance, the photo of M-104 took 10.2 hours. The deep-field photo above took over 48 hours.
And so amateur astronomers who want pictures like those taken by the professionals use CCD cameras specifically
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Astronomical%20Observation.html

5/7

1/3/2016

Introduction to Physical Astronomy - Astronomical Observation

designed to fit the telescope in place of the eyepiece. They take black and white photos of the same object three times,
through red, green and blue filters, and then composite them into a color image. And they take many photographs a second,
using computer software to scan through the tens or hundreds of thousands of images, discarding those blurred by
momentary atmospheric turbulence, and then "stack" the images (literally adding the intensities) to produce the final product.
A stroll through the Astronomy Picture of the Day Archive will show many examples of amateur photographs which rival
some of those taken professionally.

Cosmic Ray Astronomy


Cosmic rays are not radiation at all: they are high-energy atomic nuclei, mostly protons. Their energies range from 1 GeV
(109 electron volts) up to (rarely) 1020 eV. Those up to around 1015 eV are thought to originate from within our galaxy,
while the more energetic cosmic rays are thought to originate outside of the galaxy. The energy density of cosmic rays in the
galaxy is around 10 MeV per cm3, maintained largely by exploding supernovae and stellar winds. Those cosmic rays
originating outside the galaxy come from within a radius of about 250 million light years. Above approximately 4 * 1019 eV,
cosmic rays interact with the Cosmic Microwave Background Radiation; this limits the range of the most energetic cosmic
rays.
When cosmic rays hit the atmosphere they produce a shower of particles. It is this shower which is detected by cosmic ray
observatories such as the Pierre Auger Cosmic Ray Observatory in western Argentina. Their goal is not to produce an
image of a cosmic ray source, but to simply identify the sources and measure the energies associated with cosmic rays. To
do this, the Auger Observatory has constructed 1600 detectors covering an area of about 3000 km2. Events with energies
over 1019 eV have a flux of about 1 per km2 per year.

Neutrino Astronomy
The purpose of neutrino astronomy is much the same as that of cosmic ray astronomy: to detect sources and measure
energies. Neutrinos are electrically neutral, almost massless, and interact so weakly that the flux of solar neutrinos through
the Earth of approximately 65 billion per cm3 passes through the Earth each second with only a handful of interactions.
These facts makes this perhaps the most difficult astronomical undertaking besides the detection of gravitational waves.
The Kamioka Observatory in Japan has been instrumental in confirming our understanding about supernova explosions, and
in determining that neutrinos indeed have mass. Its instrument, Super-Kamiokande, is comprised of 13027 photomultiplier
tubes in 50000 tons of pure water. The tubes detect Cherenkov Radiation: electromagnetic energy emitted by charged
particles whose speed is greater than the speed of light in water (about 2.25 * 108 m/s). When a neutrino interacts with an
atom in the water, an electron or muon is created whose track the tubes can measure. The instrument typically detects less
than 14 solar neutrino events per day.
The latest high-energy neutrino observatory is currently under construction using 1 km3 of Antarctic ice: the IceCube
Neutrino Observatory. It will focus on neutrinos entering the opposite side of the Earth so that the lower-energy neutrinos
associated with cosmic ray showers will not be confused with those of extraterrestrial origin.

Gravitational Wave Astronomy


In 1974, Taylor and Hulse found a binary system of neutron stars, one of which is a pulsar. After two decades of
observation, they determined that the change in the rate of spin of the pulsar matched the predictions of General Relativity
for such a system emitting gravitational waves:
The changing orbit of binary pulsar PSR 1913 + 16 meets General Relativity. (source)

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Astronomical%20Observation.html

6/7

1/3/2016

Introduction to Physical Astronomy - Astronomical Observation

But no one has ever directly detected a gravitational wave. It is the purpose of LIGO - the Laser Interferometer
Gravitational-wave Observatory - to change that.
LIGO consists of two L-shaped detectors, each 4 km long, one in Hanford, Washington, and the other in Livingston,
Louisiana. In each, laser light travels repeatedly from one end to the other, reflected by mirrors. A passing gravitational
wave will change the relative lengths of the two beams, and the change in the interference pattern will be registered by a
photodetector. The fifth science run (recently completed) achieved a sensitivity of one part in 1021 from 70 Hz up, enabling
detection of binary inspiral of 1.4 solar mass neutron stars at a distance of 12 Mpc. No news yet on what if anything they
have seen.
There's a nice TED talk by Janna Levin which includes some auditory simulations of gravitational waves emitted by
inspiraling black holes.
Table of Contents
References
Index
2012, Kenneth R. Koehler. All Rights Reserved. This document may be freely reproduced provided that this copyright
notice is included.
Please send comments or suggestions to the author.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Astronomical%20Observation.html

7/7

1/3/2016

Introduction to Physical Astronomy - Black Holes

Black Holes
If the final mass of a collapsing star is greater than 3-3.2 solar masses, the remnant collapses behind the horizon of a black hole.
(View Cosmos DVD 6, episode 9, on Flatland and curved space.)
Note that since mass is a scalar (having no directionality), it is a constant in every frame of reference.
For a black hole of mass M and angular momentum L, the horizon is a spherical surface located at a distance
r = (G M2 + (G2 M4 - c2 L2)1/2) / (M c2)
from the center of the black hole. This is a complicated function, so we will specialize for a moment to the (probably) unphysical
case of a static black hole, which does not spin. Its horizon is then at
r = 2 G M / c2
= 2953 meters * M / Msolar
Note that this implies that G / c2 is a conversion factor for converting mass to length.
The escape velocity necessary to escape the gravitational pull of a black hole gets larger as you get closer to the horizon. Using this
expression for r, the escape velocity at the horizon is
ev = (2 G M / r)1/2
=c
which is consistent with our notions
a. of nothing (particularly light) being able to escape from inside the horizon of a black hole, and
b. all light cones are tangent to the horizon of a black hole.
The surface gravity at the horizon is
g = c4 / (4 G M)
= 1.52 * 1013 m/s2 / (M / Msolar)
For M = 2 Msolar, this is almost 2 million times the surface gravity of Sirius B, and 4 times that of the neutron star in the center of
the Crab Nebula. But there is a caveat for the neutron star: it is a pulsar, so we know it spins, and therefore using our simplistic
version of the equation for the horizon is not exactly a valid comparison. However, even though the pulsar is spinning 30 times each
second, the angular momentum term is less than 0.03% of the mass term, so this result is very close.
Consider now the tidal acceleration experienced by a 2-meter object at the horizon:
G M / r2 - G M / (r+2)2
= (c6 / (4 G M)) (c2 + 2 G M) / (c2 + G M)2
Since M is a multiple of Msolar, G M is much greater than c2, and this reduces to approximately
c6 / (2 G2 M2)
= 2 * 1010 / (M / Msolar)2
For M = 2 Msolar, this is about 5 * 109 m/s2. This means that our 2-meter object experiences a tidal acceleration over 500 million
times Earth's surface gravity! But for M = 106 Msolar, the tidal acceleration is 0.02 m/s2: unnoticeable.
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Black%20Holes.html

1/4

1/3/2016

Introduction to Physical Astronomy - Black Holes

It is expected that many if not most galaxies harbor a supermassive black hole in their cores, probably surrounded by an
accretion disc, whose matter is accelerated to relativistic speeds and radiates tremendous amounts of energy as it falls into the
horizon:
Accretion disc surrounding probable black hole in NGC 4261. (source)

Here is a movie of the crab pulsar (source). The disc-like disturbances around the pulsar and the accretion discs around the black
holes are related in that the curvature of spacetime is essentially the same (except in magnitude) around all spinning masses. It is
probable that most if not all of the violent events we see in the universe, from novae to gamma ray bursts, are powered by either
stellar collapse,
the accretion of matter onto a massive compact object like a white dwarf, neutron star or black hole, or
mergers of same.
Portfolio Exercise: Find a white dwarf other than Sirius B and a neutron star other than the one at the heart of the
Crab Nebula, and compute the following: its density, the escape velocity at its surface, and its surface gravity. For the
white dwarf, you will need to know its radius and mass; you will have to find its mass (which means it will probably
have to be a binary companion), and you can compute its radius if you find its absolute magnitude and temperature.
For the neutron star, your source will have to provide its mass and radius.
Now compute the horizon radius for a black hole of mass equal to the masses of the white dwarf and neutron star you
found. Using that radius, compute the escape velocity and surface gravity of the "equivalent" black hole. Express all
escape velocities in terms of the speed of light, and all surface gravities in terms of Earth's surface gravity (9.8 m/s2).

Curvature and Geodesics around a Black Hole


The hardest thing to conceptualize in General Relativity is how spacetime can be curved: how can something without substance
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Black%20Holes.html

2/4

1/3/2016

Introduction to Physical Astronomy - Black Holes

have shape? With enough play time around a black hole, perhaps this can become clear, or at least clearer than it is now. Our black
hole is electrically neutral, so it is described by only two parameters: its mass and its angular momentum. A collapsing star must have
a mass "M" greater than (at least) 1.86 times the mass of the Sun to form a black hole, and its angular momentum "a" must be less
than or equal to "M" for the black hole to have a horizon.
In the following applet, you can choose "M" and "a" for the black hole. You can also choose the initial position (relative to the
horizon) and speed of a pair of probes, and their initial orbital direction: either azimuthal (around the "equator") or polar (along a
"great circle" through the "north pole"). The "Replot" button starts the computations. Please be patient: each plot requires evaluation
at over 3.5 million points! The "Replot" button will be labeled with ellipsis while the evaluation process is taking place.
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows system, download the
Java VM (Virtual Machine) for your version of Windows at the download section at java.sun.com.
The plot on the left is a 3-D plot showing the curvature of spacetime around the black hole, and the paths of the probes, which
follow geodesics. The Geodesic Equations describe the path of any object whose own mass is negligible compared to the mass of
the black hole (including light rays). The equations become very "stiff" near the horizon, so the last segment or two of the geodesic
plots may not be smooth. The curvature is measured by the Kretschmann Invariant. Spacetime is described by a metric, which
literally tells how the measurement of intervals varies from event to event. The Kretschmann Invariant is an algebraic function of the
metric and its derivatives, which measures the curvature in a way that does not depend on the coordinate system used to describe
the metric.
The 3-D plot is colored magenta where the Kretschmann Invariant is positive, and cyan where it is negative. Where the plot appears
blue or white, you are seeing positive regions behind negative ones, or vise versa. The axes on the three dimensional plot are not
drawn inside the horizon of the black hole. The region inside the horizon on the 3-D plot does not appear to be completely black
because you are seeing it through some nonzero values of the invariant. The plot scale automatically changes so that the width of the
plot is four times the radius of the horizon assuming "a" is 0. To zoom, turn the scroll wheel; to change the perspective, drag the
mouse across the window.
The applet also provides a two dimensional plot (on the right) of time vs. position of the geodesics. The axes menus allow you to
control the two dimensional plot. The origin of this plot is (0, 0) except when plotting radial distance, when it is (horizon radius, 0).
The coordinates of the upper right hand corner are determined by the maximum values begin plotted (given in the "Final conditions"
window at the bottom of the applet). For the polar and azimuthal angles, the right hand edge is at /2 and 2 , respectively. The
polar plot includes a yellow line which marks the value of the polar angle where the curvature is zero at the horizon.
Things to try...
Lower the initial speed of the probes to 50% of the speed of light. This is less than the escape velocity and so the probes
cross the horizon; but when? By default, the 2-D graph shows radial distance versus proper time (which is the time the
probes measure). You can see that the probes cross the horizon in a finite amount of time, and the "Final conditions" data
indicates that it is a very short time indeed. But if you choose "asymptotic time" for the 2-D vertical axis, a very different
graph appears. We see that from the point of view of an observer "at infinity", the probes never cross the horizon but only
approach it asymptotically. This is an eloquent indication that the coordinates we use to make measurements far from the
black hole do not have the expected meaning in the region near the horizon.
You can see that changing "M" is only interesting if "a" is nonzero; this is because the applet automatically changes scale with
"M". You can also see that "M" must be small and "a" must be a significant fraction of "M" before the plot changes
substantially. This is a physical consequence of the spacetime metric around the hole.
Start the probes in a polar orbit with "a" = 1 (an extremal black hole); observe that near the horizon, in the region called the
ergosphere, the angular momentum of the black hole "drags" the probes in the direction of the black hole's rotation. This is
called frame dragging.
Note that as "a" increases, regions of "negative curvature" appear at the poles. What can this mean physically? The paths of
the polar geodesics indicate that the regions of "negative curvature" correspond to centrifugal barriers.
Vary the initial position and speed to see how the escape velocity depends on the initial position. (You have found the escape
velocity when the 2-D radial distance plot for the inner probe is a vertical line, or leans just slightly to the right.) Do this first
for a = 0 and both large and small masses. Then set a = .9 M and do the same. Does the escape velocity depend on the
mass? Does it matter whether the probes follow an azimuthal or polar orbit?
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Black%20Holes.html

3/4

1/3/2016

Introduction to Physical Astronomy - Black Holes

We found that the escape velocity at the horizon is the speed of light. Yet you may have noticed that even when
starting the probes from a distance, at almost the speed of light, sometimes they enter the horizon. While it is
true that the Newtonian escape velocity is the speed of light at the horizon, Einstein's General Relativity includes
additional effects which imply that there is a region around the black hole for which there are no circular orbits.
Since our probes start in a direction tangent to a circle around the black hole, in that region they must fall in!
We have really cheated a bit here for dramatic effect: we really don't need a black hole to see the same effects. The Kerr metric
used in this applet describes spacetime around any isolated, compact rotating massive object. So the same effects, albeit much less
pronounced, can in principle be measured around our Sun, or even around the Earth or the Moon.
Portfolio Exercise: Using the applet, vary the initial position and speed to see how the escape velocity depends on
distance from the horizon. Do this first for a = 0 and masses of 2, 2000 and 2000000 Msolar. Then set a = .9 M and
do the same. Does the escape velocity depend on the mass? Does it matter whether the probes follow an azimuthal or
polar orbit?

Falling in...
The surface of the future light cone is where a particle could be if it travels at the speed of light, and the interior of the cone is where
it could be if it travels more slowly. Therefore photons are destined to stay on the surface of their light cones and massive particles
are destined to stay in the interior of their light cones.
Contrary to what you might have heard, this means that you will indeed expire when you cross the horizon of a black hole.
Assuming you cross feet first, the future light cone of your feet is pointing into the center of the black hole, so no nerve impulses can
reach your head from there. When your heart crosses, no blood can flow from it to your head. And when your head crosses the
horizon, the parts of your brain cannot communicate with each other and your consciousness ceases.
Table of Contents
References
Index
2010, Kenneth R. Koehler. All Rights Reserved. This document may be freely reproduced provided that this copyright notice is
included.
Please send comments or suggestions to the author.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Black%20Holes.html

4/4

1/3/2016

Introduction to Physical Astronomy - Cosmic Microwave Background Radiation

Cosmic Microwave Background Radiation


After collecting seven years of data from the Wilkinson Microwave Anisotropy Probe (WMAP), the WMAP Science Team released this image of the Cosmic
Microwave Background Radiation (CMBR):
7 year data from the Wilkinson Microwave Anisotropy Probe. (source)

This image is a weighted linear combination of data in five radio frequency bands (23, 33, 41, 61 and 94 GHz) which minimizes the foreground contamination from the
Milky Way. It is a snapshot of the photosphere or last scattering surface: our view of the universe when it first became transparent to electromagnetic radiation. The
color scale in this image corresponds to temperatures of from 2.7248 to 2.7252 K (blue to red). This represents a deviation from perfect isotropy of one part in 13625.
Note that the last scattering surface depends on both the time of decoupling and the place and time of observation. The expansion of the universe changes
the distance to the last scattering surface, and the passage of time increases our horizon, so we see more of the universe at decoupling as time passes.
The minute deviations from perfect isotropy in the CMBR tells us that there were small inhomogeneities in the plasma:
dark matter density fluctuations created small "pockets" of gravitational attraction into which baryons fell;
opposing that gravitational infall was the radiation pressure, and these opposing forces caused the plasma to oscillate;
during the radiation-dominated era, the fluctuations grew with time until the length scale of the fluctuations exceeded the horizon size (the distance light can travel
during "one" fluctuation);
during the matter-dominated era, the fluctuations grew more slowly;
at decoupling, fluctuations were damped at all scales less than about 10 Mpc (so dark matter is required for galaxy formation);
the density fluctuations froze out when the universe became dominated by dark-energy.
These oscillations occurred at different length scales and left their imprint on the CMBR we observe today. Of course, those imprints have been modified by the
expanding universe, and by other factors we have yet to discuss.
By comparing the CMBR temperatures at different angles, we can construct the power spectrum: a measure of the correlations between temperature as a function of
angle.
The power spectrum which best fits the 7-year WMAP data. (source)

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Cosmic%20Microwave%20Background%20Radiation.html

1/3

1/3/2016

Introduction to Physical Astronomy - Cosmic Microwave Background Radiation

The horizontal scale is the multipole moment, denoted l; the angular scale is actually /l, so larger values of l denote smaller angles. The vertical scale is the temperature
correlation function Cl, multiplied by l(l+1)/2, and is measured in K2, or (10-6K)2. The blue region represents the range of data values within one standard deviation of
the fit line.
The peaks in the power spectrum are called acoustic peaks, and indicate the length scales of the oscillations in the plasma before the temperature fluctuations froze out
in the last scattering surface (at decoupling). The height and positioning of the peaks are influenced by a number of cosmological parameters. We will focus on the
following:
b - additional baryon mass provides additional gravitational instability and increases the amplitudes (sizes) of the acoustic peaks - if b is a larger fraction of
M, there is more inertia and the speed of sound in the plasma is smaller; this can make the sound horizon smaller (the region reachable by a given oscillation)
and therefore the acoustic peaks occur at smaller scales
CDM - increases gravitational density fluctuations
vac - changes transition to dark energy-domination; at that time, density perturbations cease growing - if vac is larger, structures spread out more
optical - after the last scattering surface freezes out (optical = 1), matter structures begin to form in the early universe; when the first stars begin burning, much of
the universe is re-ionized (from around z=10.7 until about z=6; optical < 1); this is the optical depth at (time of) re-ionization - the re-ionization smooths out
temperature fluctuations (earlier re-ionization produces more smoothing)
In addition, any parameter which affects k (primarily the s) changes the angular scales we observe. Photons of the CMBR follow geodesics; the Geodesic Equations for
the FRW metric tell us that the rate an angle changes is inversely proportional to the square of the scale factor. From the graphs above, we can see that smaller values of
k cause the power spectrum to shift to higher values of l.
Note that the length scales corresponding to the acoustic peaks should correlate with the distances between large structures in the universe. Survey results have recently
begun to confirm that expectation.

Comparing CMBFAST Results


We can use CMBFAST to explore regions of the space of possible cosmological parameters. The following applet will allow an easy comparison of the effects of various
parameter values with the current best fit to the data. When you run CMBFAST, use the default parameter values for everything except the parameter(s) you have been
assigned. Using CMBFAST, we can investigate the following:
1.
2.
3.
4.

the effects of altering the ratio of b to CDM without changing M;


the effects of changing the ratio of vac to M;
the effects of making > 1 or < 1 without changing the ratio of b to CDM to vac, and
the effects of varying the optical depth.

After you run CMBFAST (which may take a few minutes), your results will be returned to you. Look for the number on the "Output Files" (ie., "02036736" in
"cmb_02036736.fcl"). Enter that number in the applet below and it will load the power spectrum data and plot it. Each time you enter a different number, it will be
plotted in a different color (from red to violet). The best fit plot is plotted in black.
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows system, download the Java VM (Virtual Machine) for your
version of Windows at the download section at java.sun.com.
We have ignored changes in the temperature anisotropy due to gravitational lensing. We have also ignored ionization from early hot clusters, and gravitational red shifting
from the galaxy foreground (between us and the last scattering surface). Finally, all of our deliberations are strictly with the temperature anisotropies; there are also
polarization anisotropies: deviations in the electric field and magnetic field directions in the CMBR. These promise to be important sources of information in future studies.
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Cosmic%20Microwave%20Background%20Radiation.html

2/3

1/3/2016

Introduction to Physical Astronomy - Cosmic Microwave Background Radiation

Portfolio Exercise: Using CMBFAST and the comparison applet, perform the 4 investigations listed above the applet. Do at least 4 trials for each
investigation. Be sure that Omegab is greater than 0 for each trial.
Capture images of each of the comparison plots for the 4 investigations, and annotate them, describing which plots belong to which parameter values.
What is the relationship between the results of investigation number 3 and the geometry of the universe?
Include all 4 in your portfolio.
The Planck Collaboration has recently released data from their first 15.5 months which estimate the Hubble parameter (H0) at 67.15
km/s/Mpc, b at 0.049, CDM at .268 and vac at .683. Run CMBFAST with this profile and compare it to the others. Include it in your
portfolio.

An Overview of Large-scale Structure Formation


This subject is relatively young, but we can build up what seems to be a reasonable picture as follows. Think of the formation of large-scale structures in the universe
(galactic clusters, voids, etc.) as being a result of the merging of dark matter halos:
expansion of the universe causes under-dense regions to grow, forming voids;
voids are separated by sheets;
filaments form where sheets meet;
nodes form where filaments intersect;
the nodes correspond to cluster halos (interesting note: these halos tend to be triaxial ellipsoids);
as the universe continues to expand, the process builds structure from the bottom up: halos merge, accrete.
At re-ionization, the speed of sound was about 10 km/s, so halos with escape speeds less than that (around 108 Msolar) didn't form stars (the gas would not fall into the
halos).
Star formation drove gas out of the galactic halos (accounting for the presence of heavy elements in the intracluster medium); this process was aided by ram pressure as
the galaxies moved through the intracluster medium.
Table of Contents
References
Index
2010, Kenneth R. Koehler. All Rights Reserved. This document may be freely reproduced provided that this copyright notice is included.
Please send comments or suggestions to the author.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Cosmic%20Microwave%20Background%20Radiation.html

3/3

1/3/2016

Introduction to Physical Astronomy - Cosmology

Cosmology
History of the Universe
The average temperature of the universe is inversely related to the scale factor:
temperaturepast / temperaturenow = scale factornow / scale factorpast
= red shift (z) + 1
So as the universe expanded, the temperature dropped. We know that each particle has an energy associated with it through the equation
energy = mass * c2
Similarly, each of the fundamental forces has an energy scale associated with it. By dividing each energy by Boltzmann's constant (1.381 * 10-23
J/K):
temperature freeze out = energy / kB
we find the temperature at which each force becomes effective, and at which each particle condenses from its constituent parts. For instance, above
1015 K, the particles which mediate the weak force (W and Z bosons) are unstable to pair production (decay into an electron and a positron), and
interactions which change quark flavors (as in radioactive decay) are not possible. When the temperature drops below that associated with the mass of
the proton:
1.673 * 10-27 kg * c2 / kB
= 1013 K
the constituent quarks in the proton are cool enough to be bound together by the strong force:
event

scale factornow / scale factorthen

temperature (K)

time

strong forces freeze out

1027

3.7 * 1026

10-35 s

weak forces freeze out

1015

3.7 * 1014

10-10 s

protons, neutrons freeze out

1013

3.7 * 1012

0.0001 s

1010

1010

neutrinos decouple

3*

electrons freeze out

6 * 109

2.2 * 109

100 s

primordial 2H, 4He form

9 * 108

3.3 * 108

2-15 minutes

1.1 *

1s

When the protons and neutrons froze out, the ratio of protons to neutrons was about 6:1 because of the mass difference between the
neutron and proton (the neutron is slightly heavier). The large number of neutrons available, as well as the fact that neutron capture occurs
faster than proton fusion, caused the nucleosynthesis reactions here to be somewhat different from those taking place in the Sun. Once
the universe cooled enough to allow Deuterium (2H) to exist, the ratio was 7:1 (from neutron decays) and the following reactions
occurred:
1H + n -> 2H +
2H + 2H -> 3He + n +
3He + n -> 4He +

This sequence lasted about 15 minutes. The final ratio of 4He to 1H was about 1:12, so that the universe was about 75% Hydrogen and
25% Helium by mass. A small amount of Deuterium survived; since it does not survive in stars, what Deuterium we observe is
primordial. This is a sensitive indicator of the density of normal matter (not dark) in the universe, since a denser universe would have
contained more protons and produced more Deuterium during nucleosynthesis.
Heavier elements were not formed because the temperature and density were both dropping very quickly; in stars they do form because
the temperature and density increase (slowly).
We now take up our history:
event

scale factornow / scale factorthen

temperature (K)

time

photons decouple, atoms form

3000

1091

377000 years

first stars

60

10.4

109 years

today

2.73

1.378 * 1010 years

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Cosmology.html

1/9

1/3/2016

Introduction to Physical Astronomy - Cosmology

Decoupling means that those particles are no longer in thermal equilibrium with their environment. When neutrinos decoupled, the
universe became transparent to them; similarly for photons.
The scale factors in the tables above were obtained by using Tnow = 2.725 K, the temperature of the Cosmic Microwave Background Radiation
(CMBR). Study of the CMBR has made cosmology an experimental science, as we shall soon see.
The temperatures in our history indicate that the early universe was filled with intense radiation. Even before the decoupling of photons, the density of
matter became greater than the density of radiation (due to red-shifting, as we shall see). Wide field surveys indicate that the universe is now very close
to being a homogeneous "dust" of galaxy clusters. The pressure from the radiation has all but vanished. But the Type IA supernovae surveys we
discussed previously indicate that the expansion of the universe is accelerating. We do not know what causes this acceleration, and for now we simply
give it the commonly-used name dark energy. So the universe has passed through three distinct phases:
1. radiation-dominated phase (z > 3250),
2. matter-dominated phase (3250 > z > 0.37), and now
3. dark energy-dominated phase (z < 0.37):
Log-Log plot of temperature vs scale factor.

General Relativity
General Relativity (GR), which relates the geometric qualities of spacetime to the matter and energy it is filled with, provides us with a mathematical
context for understanding the evolution of the universe. Spacetime is described by a metric: a rule for how to measure intervals. The usual procedure in
GR is to find the most general metric which is consistent with the symmetries of the problem at hand, and to find the most general form of the
expressions describing the matter and energy. These are related by Einstein's Equations, which are then solved for relations between the free
parameters.
Given the wide field surveys and observations of the CMBR, we need the most general metric which is homogeneous and isotropic (the same in all
directions). This is the Friedmann-Robertson-Walker (FRW) metric, and it is described by two parameters. The first is "k", the curvature
constant; if we choose a time "t" and take all the points in the spacetime which have the same value of t (called a spatial section),
k = 1 means the section is positively curved:
cross sections in a fixed direction are circles;
the universe is closed like a sphere;
the sum of the interior angles in a triangle is more than 180 degrees;
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Cosmology.html

2/9

1/3/2016

Introduction to Physical Astronomy - Cosmology

k = 0 means the section is not curved; it is flat space:


cross sections in a fixed direction are straight lines;
the sum of the interior angles in a triangle is 180 degrees;
k = -1 means the section is negatively curved:
cross sections in a fixed direction are hyperbolas;
the neighborhood of every point in the universe is shaped like a saddle;
the sum of the interior angles in a triangle is less than 180 degrees:
Surfaces of constant curvature, and their triangles.

(Strictly speaking, what we have said about k, and the images above, are relevant to 2-dimensional surfaces. You have to use your
imagination to see how all this applies to 3 dimensions, but with a little practice it can be done.)
For nonpositive k, the universe can be either finite (although obviously very large!) or infinite. A finite universe need not have an edge: a flat
closed universe (finite, without boundary) could be like a cube, but with identifications: walking "out" one side is the same as walking "in" the
opposite side. Similarly, a negatively curved space can have identifications, but these are much harder to visualize. An infinite universe is called
an open universe (without boundary). If k is positive, the universe cannot be open.
The other parameter is a function "a(t)"; this is the scale factor we mentioned earlier. It measures the "size" of the universe as a function of time. The
Big Bang occurred when a(t) was zero, and the expansion of the universe means that a(t) increases as a function of time. The Hubble Parameter is
defined as the rate of change of a(t) divided by a(t). The rate of change of a(t) is denoted a'(t), and it must be positive as long as the universe is
expanding:
H(t) = a'(t) / a(t)
Our matter/energy expression must be able to describe our three act history of the universe: radiation-dominated, matter-dominated and dark energydominated. This can be done using a perfect fluid, which is described by its energy density "" and pressure "p" (perfect fluids do not have viscosity
or convection). We will usually assume a simple (but reasonable) equation of state (which gives the pressure as a function of the density)
p=w
where "w" is a constant:
if w=0, the equation describes dust;
if w=1/3 it describes radiation, and
if w=-1 it describes the Cosmological Constant.
If the dark energy is not due to a Cosmological Constant, it is usually given the name quintessence, and it has a different equation of state. Current
data is consistent with a Cosmological Constant, and we will assume that simple scenario in the following.
With these parameterizations, Einstein's Equations reduce to two simple equations:
a'2(t) ((t) / c - 1) = k c2,
and
a''(t) = - (4 G / 3 c2) a(t) (3 p(t) + (t))
= - (3 w + 1) (4 G / 3 c2) a(t) (t)
where we have shown explicitly the dependence of the density and pressure on the time (since density and pressure both depend on volume and
therefore on the scale factor, which is a function of time). c is the critical density
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Cosmology.html

3/9

1/3/2016

Introduction to Physical Astronomy - Cosmology

3 H2 c2 / 8 G
= 8.2487 * 10-10 J / m3,
equivalent to about 5.5 protons per cubic meter.
If = c, k must equal zero.
If > c, k must be positive, and
if < c, k must be negative.
Note that unless w < -1/3, the second of Einstein's Equations implies that the acceleration of the universe must slow down (since a''(t)<0).
Einstein's Equations are supplemented by a conservation equation which guarantees that energy is conserved. With our simple equation of state, the
conservation equation has the solution
(t) is proportional to 1 / a3(w + 1)(t).
This tells us something we already knew: for matter, the density drops as the volume increases, and the radiation pressure drops as a4(t) increases, due
to the combination of the increase in volume and the red shift.
Note that for the Cosmological Constant, w = -1, so is a constant: it does not depend on a or t.
It is customary to use the ratio / c as a cosmological parameter. This quantity is denoted "". Using it, our first Einstein Equation (also called the
Friedmann Equation) is written
a'2(t) ( - 1) = k c2
So if = 1, k must be zero; if < 1, k must equal -1, and if > 1, k must equal 1.
When the universe cooled so much that there was insufficient energy to ionize Hydrogen atoms, the universe became transparent to photons. Before
then, it consisted of a dense plasma (electrically charged fluid) containing electrons, baryons, and photons. There was also dark matter: massive
particles (not yet understood as part of the Standard Model) which do not exchange photons, and so do not interact electromagnetically. Dark matter
may come in two forms: Cold Dark Matter (CDM) or Hot Dark Matter (HDM); HDM travels at speeds close to that of light, while CDM is
non-relativistic (speeds 50 km/s). There was also dark energy, but its influence appears to have been negligible in the early universe.
Since the baryons are roughly 2000 times more massive than electrons, they were the fundamental source of inertia (resistance to acceleration).
Baryons and dark matter were the fundamental sources of gravitational attraction.
Photons were the fundamental source of pressure, which was therefore experienced by the electrons and baryons but not the dark matter.
Neutrinos contributed to the radiation energy density because they are nearly massless and move at relativistic speeds, but because they are
electrically neutral they did not contribute to the pressure: they simply passed through the plasma unimpeded.
Because of their differing physical effects, the contributions of radiation, baryons, HDM, CDM and the dark energy to are distinguished by
subscripts:
= rad + b + HDM + CDM + vac
where "vac" denotes the dark energy contribution. Sometimes "" is used instead of "vac" when we are particularly interested in the cosmological
constant. In addition, we sometimes write
M = b + HDM + CDM
to denote the contribution of matter to .

Solving the Friedmann Equation


With a choice of k and w, we can solve the Friedmann Equation for radiation, matter or a Cosmological Constant:
Solutions of the Friedmann Equation for radiation (in red), dust (in green) and Cosmological Constant (in blue), for each value of k.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Cosmology.html

4/9

1/3/2016

Introduction to Physical Astronomy - Cosmology

The graph on the right shows the solutions for a Cosmological Constant; these solutions are exponential and must be fitted to the appropriate solution
from the graph on the left at the appropriate time. From these graphs we see that for smaller values of k, the universe expands more rapidly, and that
radiation tends to make the universe expand more rapidly than dust.
The universe of course has all three components: matter, radiation and Cosmological Constant. When all three are present, it is quite difficult to solve
the Friedmann Equation analytically, but it can be solved numerically. In order to do so, we must have values for the various s and k.
While our equation of state above implies that we can have either matter or radiation, but not both, the solutions to the conservation equation are valid
even when both are present. Equating the energy densities as functions of the scale factor allows us to find the red shift at the transition from the
radiation-dominated era to the matter-dominated era:
radiation matter
rad / (aradM)4 = M / (aradM)3
aradM = rad / M
zradM = M / rad - 1
Using M = 0.278, we can invert the equation to compute rad = 8.55 * 10-5, which includes contributions from both photons and neutrinos.
The energy density from photons alone can be found from the black body distribution as
= 8 5 (kB * TCMBR)4 / (15 (h c)3)
= 5.06 * 10-5 c
We can perform a similar calculation, using matter, to find &Lambda.
Portfolio Exercise: Verify the computations of rad and , and compute &Lambda. For &Lambda, use zM = 0.374397.
For our numerical solution, we have assumed values for H0, zradM, M, and k consistent with the WMAP7+BAO+SNSALT data set for the
CDM model. The numerical solution for the scale factor is shown below in blue; the left plot shows the past, and the right plot shows the future:
The scale factor up to the present (left) and in the future (right).

Up to the present day, the numerical scale factor is approximately


apast(t) = (t / t now)2/3
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Cosmology.html

5/9

1/3/2016

Introduction to Physical Astronomy - Cosmology

(left, in red), but in the distant future it increases exponentially:


afuture(t) = 4/3 ec t ( / 3)
(right, in red). apast(t) and afuture(t) are the solutions for the matter-dominated era and the Cosmological Constant-dominated eras, respectively. Since
they are in such good agreement with the complete numerical solution, we will use them in the following. Of course they do not quite match up, and
they neglect the fact that in the recent past, a''(t) becomes positive, as it must since we observe the expansion of the universe to be accelerating. For
those shortcomings, they still are very good approximations, and will illustrate all of the qualitative points we wish to make.
There's a caveat here: for the past, we are using the solution for the matter-dominated universe, and in the early stages of its evolution, the
universe was dominated by radiation. The scale factor was essentially linear in t during that era, but we can't match the solutions easily
because we have no distance information from those early times. So we will pretend this value is accurate, but remember that this is only
an approximation.
Recalling our discussion of the Hubble relation, we know that the expansion speed of an object located at a distance r(t) is
vexp = H(t) * r(t)
= a'(t) * r(t) / a(t)
= a'(t) * r(tnow),
since r(t) = r(tnow) * a(t). Using apast(t) for a(t), we see that the Hubble parameter changes with time up to the present as
H(a) = 2 / (3 tnow a3/2)
The Hubble parameter / H0, from past to present as a function of scale factor.

But using a(t) = afuture(t) implies that in the future, the Hubble parameter approaches a constant: c (/3), or approximately 11.85 km/s / Mpc.

We have seen images of proto-galaxies at z = 1 and beyond. Let us assume that an extraterrestrial named Bob lives on an early-developed planet in
one of those protogalaxies (at z=1), and is not moving with respect to it. Using apast(t), it is possible to compute the current distance to Bob, even
though the light he emits now will not be seen by us for a long time. Our equation for vexp is actually a differential equation:
dr(t)/dt = a'(t)/a(t) * r(t)
If we consider a light ray traveling toward us from the past, this equation describes the path of the ray as affected by the expansion of the universe:
dr(t)/dt = a'past(t)/apast(t) * r(t) - c
The expansion of the universe essentially modifies the speed of light, which is only guaranteed to be constant in inertial frames: coordinate systems
moving at constant velocity. The solution to this equation is
rpast(t) = bpast * t2/3 - 3 c t,
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Cosmology.html

6/9

1/3/2016

Introduction to Physical Astronomy - Cosmology

where bpast is a constant to be determined by setting rpast(tnow) = 0, since we are measuring distances relative to our current position (where we see
the light from Bob).
Since Bob was at z = 1 when he emitted light that we are just seeing now at z = 0 (a = 1), we can find the time since the emission by setting
apast(tthen) = 1 / (z + 1)
and solving for tthen. Using the current age of the universe for tnow we can find bpast, and therefore Bob's distance when he emitted the light
(rpast(tthen)). By multiplying by anow/apast, we then obtain his current distance rnow from us. We find that Bob is currently 1.146 * 1026, or 3.71 Gpc,
away.
With Bob's current distance from us, we can now compute Bob's distance from us at any time, and his recession velocity at any time as well. Here are
plots of his recession velocity (with c = 1), in the past (on the left) and in the future (on the right). The horizontal scale is the scale factor, and the red
line corresponds to Bob, while the blue line corresponds to an object midway between us:
Expansion speeds of Bob (red) and an object between us (blue), vs scale factor (c = 1).

We can see that Bob's recession velocity is greater than the speed of light for a < 0.26 (in the past) and a > 6.8 (in the future). We also see that the
recession velocity of objects closer than Bob dropped below the speed of light earlier than Bob's did, and will become greater than the speed of light
later than Bob's will. The opposite is true of objects further from us than Bob; the length of time an object's recession velocity is below the speed
of light varies inversely with its distance from us.
None of this contradicts what we have said previously about nothing being able to move faster than the speed of light. Bob is at rest: his
velocity relative to his neighbors is zero. The distance between events is increasing faster than the speed of light because the universe is
expanding there at that rate.
If we plot Bob's position at various times in his history (he is a very long-lived extraterrestrial), we see that before the scale factor was 0.26 (left), his
world line is spacelike, and after that it is timelike. It becomes spacelike again after the scale factor becomes greater than 6.8 (right):
r(t) for Bob; the black line is the edge of the light cone (past on the left, future on the right).

Once we have rnow, we can find the path of a light ray from Bob at any time by setting rpast(tthen) equal to rnow*athen, and solving for a new bpast.
There will be a new bpast for every light ray emitted by Bob at a different time.
If we consider a light ray traveling toward us from Bob at some time in the future, the solution is
rfuture(t) = bfuture * ec t ( / 3) + (3 / &Lambda),
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Cosmology.html

7/9

1/3/2016

Introduction to Physical Astronomy - Cosmology

where bfuture is a different constant to be determined by setting rfuture(tthen) = rnow*athen. (To find tthen, we set afuture(tthen) equal to athen and
solve for tthen.)
These computations help us to realize what it means to study General Relativity!
Bob's light cones are defined by a "left edge" and a "right edge". The left edge is just rpast(t) (or rfuture(t)) that we found above. The right edge is found
using
dr(t)/dt = a'past(t)/apast(t) * r(t) + c,
with new solutions
rpast(t) = bpast * t2/3 + 3 c t
and
rfuture(t) = bfuture * ec t ( / 3) - (3 / &Lambda).
Note that these equations have different values for bpast and bfuture, but those values are found the same way as before, since both "edges" of the light
cone start at the same point.
We can solve these equations for the light rays passing through any event on Bob's world line, and so reconstruct his light cones at several interesting
times:
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows system, download the Java VM (Virtual
Machine) for your version of Windows at the download section at java.sun.com.
Notice first that these light cones are not at 45 degree angles: they are warped by the expansion of the universe, whose expansion speeds depend on
position. The red dots outline Bob's world line, and the ends of the axes correspond to a = 1/101, a = 1/11, a = 1 and a = 10 (except for the last
plot). The horizontal black line on the future plots (a > 0) represents the present.
As we discussed previously, the blue area below an event on a world line is the past light cone of that event, and the blue area above it is its future light
cone. The world line of events taking place at the origin is simply the vertical axis, on the left edge of each frame. The green regions are where light
cones from our world line and Bob's overlap. A green region below an event indicates a point in our mutual causal past: a prior event there is visible
from the two later events. A green region above an event indicates a point in our mutual causal future: an event at one will eventually be visible there.
The horizontal axis represents the time of the Big Bang; at that time, the entire universe is the point at the origin.
It appears that Bob's events at a = 1/1001 are causally disconnected from those at the origin. But we must be cautious about over-interpreting this
graph; the resolution is insufficient to determine whether the light cones overlap at very early times. However, by a = 1/11 it is clear that events at both
locations are causally connected: they have a common past.
Just after a = 1/4, the incoming light ray from Bob tilts to the left of vertical: Bob's recession speed has dropped below that of light. Because we see
Bob currently at a = 1/2, that plot shows the incoming ray arriving at the origin at the top of the frame: the present. Just before a = 7, the incoming light
ray from Bob tilts back to the right of vertical as Bob is again receeding from us faster than light. But by the time the universe is 15 times its current
size, Bob's light cone will no longer share a future with ours.
In general, the further away something is, the later we first see it, and the sooner it will move beyond our particle horizon. We do not know if anything
exists outside our current particle horizon, but depending on how large the universe actually is, there may be portions that we will never see. Of
course nothing is "visible" until after a = 1/11: before then, there were no stars, and the universe was filled with an infrared glow that eventually red
shifted to become the CMBR.
Putting all this together, we see the universe expanding so rapidly during its early history that more distant regions evolved independently. In the future,
the accelerating expansion will move the galactic clusters out of each others' future light cones, and each will eventually grow cold surrounded only by
the void of the then much colder CMBR. This of course will not happen until the universe is many times older than it is now.
Perhaps we should take some consolation in realizing that we live at a point in the life of the universe when it is possible to understand its evolution. In
the far future, these notions would not ever be testable; much less would it occur to us to test them.
A Small Detail...
These graphs place us at rest at the origin; but what do we mean by the word "origin" if the universe is homogeneous and isotropic? If the universe
looks the same from every position, where is the center?
When we solved Einstein's Equations, we used co-moving coordinates: coordinates which move with the point in question. So in our
solution, any point in the universe could just as well be the origin, and the horizontal distance in the light-cone diagrams is simply the
separation between two points which are "riding along" with the expansion of spacetime.
Note that it is only on the scale of clusters of galaxies that the universe appears homogeneous and isotropic; clearly, the
universe looks different to us on Earth than it does if we were, say, on the moon. So when we say that spacetime is
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Cosmology.html

8/9

1/3/2016

Introduction to Physical Astronomy - Cosmology

expanding, we mean that the distances between galactic clusters are growing with the scale factor, not that the distance
between your nose and your eyes is changing. It might be, but the Friedmann Equation has nothing to say about that: it
simply does not apply to scales smaller than megaparsecs. If the distance between your nose and your eyes were changing,
and it is, for example, 1 cm now, the current value of the Hubble Parameter tells us that the rate of separation would be
1 cm * 69.9 km/s / Mpc
= 10-2 m * 2.265 * 10-18 / s
= 2.265 * 10-20 m/s,
or about 1/100,000th of the diameter of an atomic nucleus each second. In your lifetime, the expansion of the universe
would change that distance about the radius of an atom. So even if the Friedmann Equation did imply that the distance
between atoms is changing with the expansion of the universe (which it does not!), it would be undetectable on such small
scales.
Table of Contents
References
Index
2014, Kenneth R. Koehler. All Rights Reserved. This document may be freely reproduced provided that this copyright notice is included.
Please send comments or suggestions to the author.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Cosmology.html

9/9

1/3/2016

Introduction to Physical Astronomy - Dark Matter

Dark Matter
Galactic Rotation Velocities
Returning now to the case of UGC 2034, the Hubble Relation gives us a distance of
d = v / H0
= 572 km / s / (69.9 km/s / Mpc)
= 8.18 Mpc
We should view this as a correction to the absolute magnitude, but for the sake of simplicity we will continue to use
the value of M from the table.
We can compute the maximum speed of UGC 2034's rotation from the relation
line width = 2 * speed * central wavelength / c
by solving for the rotational speed (and including the Doppler Shift!) as
v = ((c / 2) * width / central) / (1 + vrecession / c)
= ((c / 2) * 2.43286 / 6562.78) / (1 + 572000 / c)
= 55500 m / s
The absolute magnitude can be used (as above) to determine the luminosity as
L = 10(4.83 - (-17)) / 2.5
= 5.4 * 108 Lsolar
From our statistical discussion in the portfolio exercise, the average mass to luminosity ratio in solar units is around
1/2. Hence the mass of UGC 2034 should be in the neighborhood of 2.7 * 108 Msolar. But we can use Newton's
Laws to check this. First, the radius of UGC 2034 can be computed from the angular diameter (converting from arc
minutes to radians) as
r = d * diameterangular * / (60 * 180) / 2
= 8.18 Mpc * 2.818 * / (60 * 180) / 2
= 3353 pc
Then we can solve the relation for orbital velocity
orbital velocity = (G * mass / radius)1/2
for mass, obtaining
m = v2 * r / G
= 555002 * (3353 * 3.086 * 1016) / G
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Dark%20Matter.html

1/3

1/3/2016

Introduction to Physical Astronomy - Dark Matter

= 4.776 * 1039 kg
= 2.4 * 109 Msolar
This is almost 9 times our estimate from the luminosity. We have probably over-estimated the mass because our
rotational velocity represents the maximum possible consistent with the observations. But even if our velocity is
twice the actual (which is unlikely), there is still more than twice as much mass in UGC 2034 as we can see.
While there is significant variation in the mass to luminosity ratio depending on both size and morphology, the
general conclusions are the same: assuming that Newton's Laws hold throughout the universe, it is clear that galaxies
have more matter than is visible to our telescopes!
A plot of the measured rotation curves for a sample of Sb and Sc galaxies. (source)

This issue raises an important question about the Tully-Fisher Relation between rotational velocities and luminosity:
how is the dark matter, which contributes to the rotational velocities but not to the luminosity, linked to the
luminous matter?
Portfolio Exercise: Create a table with the Hubble Distance, Doppler rotational speed, luminosity,
luminous mass, angular diameter and total mass (from the orbital velocty relation) for all 20 galaxies in
this sample.
In general, dark halos extend out to about 300 Kpc and contain roughly 10 times the stellar mass.
Some ellipticals seems to lack dark halos.
21 cm Hydrogen emission seems to track the dark halo.
Most galactic halo dark matter outside of about 50 Kpc is redistributed into cluster dark matter during
interactions.
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Dark%20Matter.html

2/3

1/3/2016

Introduction to Physical Astronomy - Dark Matter

This applet lets you "design" a galaxy and view the corresponding rotation curve (velocity as a function of distance
from the center). It assumes that within the galactic disc, the mass is evenly distributed by volume. The same
assumptions are made for the central bulge and the dark matter halo, which is drawn in violet. The disc thickness is
assumed to be 1% of its radius, and the bulge radius is limited to 50% of the disc radius. The width of the graph is
twice the largest chosen radius, and the color intensity is scaled so that the largest mass selected corresponds to the
brightest colors.
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows system,
download the Java VM (Virtual Machine) for your version of Windows at the download section at java.sun.com.
Portfolio Exercise: In the sample galaxies we have been working with, the paper lists all as spirals
except 2034, 5829, 5935 and 12060. For three of those spiral galaxies, use the applet above to
create dark matter profiles. Start with the values you have calculated for the luminous mass, disc
radius and dark matter mass. Use SIMBAD and the Aladin viewer to obtain an image of each galaxy,
and estimate the bulge radius and the percentage of luminous matter in the bulge using the image. If a
blue ("O" filter) JPEG is available, it is often the clearest image. If you cannot decide from the image,
you may assume that the bulge radius is 20 to 30 % of the disc radius, and that the percentage of
luminous mass in the bulge is about 30.
Begin with a dark matter radius of about 3 times the disc radius. Then vary the dark matter mass and
radius until the rotational velocity curve (across the top) looks approximately like the ones in the graph
above (or like the curve shown when the applet starts, which is approximately what is expected for the
Milky Way). In most cases you will have to lower the dark mass from your calculated value to
reproduce the expected velocity curve. Include images of your final profiles in your portfolio.

Cluster Dark Matter


Since galactic rotation curves first indicated the presence of dark matter, gravitational lensing has enabled
astronomers to image dark matter:
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows system,
download the Java VM (Virtual Machine) for your version of Windows at the download section at java.sun.com.
In the Bullet Cluster, gas is glowing red (in x-rays), showing the shock waves from the colliding clusters, while their
dark matter (in blue) is passing through each other relatively unimpeded.
Table of Contents
References
Index
2010, Kenneth R. Koehler. All Rights Reserved. This document may be freely reproduced provided that this
copyright notice is included.
Please send comments or suggestions to the author.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Dark%20Matter.html

3/3

1/3/2016

Introduction to Physical Astronomy - Electromagnetic Waves

Electromagnetic Waves
The light you see,
the infrared (literally, "below red") radiation that warms you,
the ultraviolet (literally, "beyond violet") rays that sunburn you,
the radio waves that carry your music and television signals and the
x-rays used to diagnose medical conditions,
are all forms of electromagnetic radiation. They are all waves: periodic functions of space and time. The length of a spatial period
is called the wavelength and is denoted . The inverse of the temporal period is called the frequency (denoted ) and is measured
in cycles per second, or Hertz (Hz). All of these waves propagate at the speed of light (c), and for our purposes, obey the simple
dispersion relation
c = wavelength * frequency
This applet might help you to better visualize waves. You can choose sine waves or cosine waves; the number on the left is just 2,
and shows you the scale of the plot. If you think of the horizontal axis as distance, then you are seeing wavelengths. If you think of it
as time, you are seeing periods. The wave number is the number of waves that fit in a given length or time, and the phase angle
determines what happens at the origin. As you can see, higher wave numbers give you smaller wavelengths and higher frequencies:
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows system, download the
Java VM (Virtual Machine) for your version of Windows at the download section at java.sun.com.
All of these types of electromagnetic radiation consist of changing electric and magnetic fields, and differ only by their wavelength (or
frequency):
radiation

(Hz)

energy (eV)

source

radio

>1m

< 3 * 108

< 1.24 * 10-6

low-energy atomic or molecular motions

microwave

> .1 mm

< 3 * 1012

< .0124

rigid molecular motions

infrared

> 7000 Angstroms

< 4.3 * 1014

< 1.78

molecular bond motions

visible light

> 4000 Angstroms

< 7.5 * 1014

< 3.1

atomic electron transitions

ultraviolet

> 50 Angstroms

< 6 * 1016

< 248

atomic electron transitions

x-rays

> .03 Angstroms

< 1020

< 414 K

electron transitions in heavy atoms

gamma rays

< .03 Angstroms

> 1020

> 414 K

nuclear decays

An Angstrom (A) is 10-10 meters. An electron-volt (eV) is 1.602 * 10-19 Joules (a 60 Watt light bulb uses 60 Joules of energy
every second).
The energy column corresponds to the energy of a single photon: the smallest amount of radiation possible. The energy (in eV) of a
photon of frequency is
energyeV = h * frequency / e,
where h is Planck's Constant (6.626 * 10-34 Js) and e is the number of Joules in an eV.
The terms "near" and "far", when applied to ultraviolet and infrared wavelengths, indicate how close they are to the visible
wavelengths.
Photons from the near-infrared to x-rays are produced when an atomic electron moves from one energy level to a lower one. The
characteristic red color seen in many nebulae is due to the Hydrogen- transition of an electron from the third to the second energy
level, with emission of a 6563 A photon:

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Electromagnetic%20Waves.html

1/4

1/3/2016

Introduction to Physical Astronomy - Electromagnetic Waves

Infrared radiation of lower energy results from internal molecular vibrations (in these examples, of triatomic molecules):

Observation in multiple wavelengths


What do these other wavelengths tell us? Consider these images ("R", "G" and "B" stand for Red, Green and Blue, respectively):
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows system, download the
Java VM (Virtual Machine) for your version of Windows at the download section at java.sun.com.
Near-Infrared observation allows us to "see through" cold dust to the warmer dust and stars inside.
Far-Infrared observation allows us to image cooler dust (as in the last LMC image).
Ultraviolet observation reveals hot, massive stars.
X-ray observations shows hot gases, as well as matter accreting onto neutron stars or black holes.
Radio observations reveal neutral Hydrogen atoms, both in the galaxy itself and in the surrounding halo.
We see that higher frequency or energy, and smaller wavelengths, correlate with higher temperature. This is natural, because
temperature is simply a measure of thermal energy.
Almost all that we know about the universe comes from these five attributes of electromagnetic radiation:
1. the direction it comes from;
Pressure waves in the Sun, caused by violent events on the surface, can travel through the Solar interior. As they
do so, they undergo refraction: their direction of propagation changes because the speed of sound changes
with depth (the speed of sound depends on density and temperature, which increases with pressure).
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Electromagnetic%20Waves.html

2/4

1/3/2016

Introduction to Physical Astronomy - Electromagnetic Waves

Astronomers have exploited these changes of direction to explore the details of the Solar interior; this is known
as Helioseismology.
2. the intensity we receive (power, or energy per unit time, over area);
The total power output over all wavelengths is called the luminosity, and is related to the intensity by
intensity = luminosity / (4 distance2)
This is often called the inverse square law, and is a result of the fact that in three spatial dimensions ("3D"), the
area of a sphere is 4 r2: the electromagnetic radiation propagates symmetrically in all directions, so surfaces of
equal intensity are spheres:
The inverse square law.

So the luminosity is determined by the source, but the intensity also depends on where you are when you are
looking at it.
3. the wavelength (or frequency);
4. the relative phase (see interferometry); and
5. the polarization (the direction of its electric field).
In general, electromagnetic waves are radiated such that their electric field directions are randomly distributed.
If, however, they pass through a medium whose molecules have been oriented along a magnetic field, the
medium can polarize the waves so that their electric fields are aligned along a single direction (this is the same
principle used in polarized sunglasses, except there it is not a magnetic field but a crystalline structure that is
responsible for the polarization). Astronomers have very cleverly used the amount of polarization to measure the
magnetic fields of the interstellar medium. Most such measurements are in radio wavelengths, and typically give
magnetic field strengths on the order of 10-5 to 10-4 that of the Earth, although fields near the galactic center
have been measured to be of order 10 times stronger.
While electromagnetic wave direction, intensity, wavelength, phase and polarization account for the vast majority of our knowledge
of the universe, there are three notable exceptions: cosmic rays, neutrinos and gravitational waves.
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Electromagnetic%20Waves.html

3/4

1/3/2016

Introduction to Physical Astronomy - Electromagnetic Waves

We often measure the spectrum of an object: the distribution of the intensity (energy per unit area per unit time) of the
electromagnetic radiation we observe as a function of wavelength. We will return to this subject soon.

Interferometry
This applet illustrates how individual waves can reinforce and cancel each other out. Press the "Start" button, turn the blue and red
waves "on", and choose the black and blue wave modes to be the same. Then vary the blue phase angle from 0 (reinforcement) to
(cancellation):
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows system, download the
Java VM (Virtual Machine) for your version of Windows at the download section at java.sun.com.
This reinforcement and cancellation varies with time and position to produce an interference pattern. This is the interference
pattern produced by two sources (which are located at the bottom of the plot):
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows system, download the
Java VM (Virtual Machine) for your version of Windows at the download section at java.sun.com.
Alternatively, if we place two or more telescopes at different locations and look at the same object, we can use the information
about how their images interfere to construct a more detailed picture of the object. The positions of the telescopes must be
determined very precisely, so this technique is primarily practical for ground-based astronomy. Since the atmosphere is opaque to
far ultraviolet and many infrared wavelengths (due to absorption by water, Oxygen and Carbon Dioxide), interferometry is limited to
optical, radio and a few near-infrared ranges.
Table of Contents
References
Index
2013, Kenneth R. Koehler. All Rights Reserved. This document may be freely reproduced provided that this copyright notice is
included.
Please send comments or suggestions to the author.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Electromagnetic%20Waves.html

4/4

1/3/2016

Introduction to Physical Astronomy - Elementary Particles

Elementary Particles
Particle physics may seem far removed from astronomy, but we will need some basic comprehension of elementary
particles before we can understand the nuclear reactions that power the stars, as well as stellar evolution, and even
cosmology. Our current knowledge of elementary particles is contained in the Standard Model of Particle Physics,
and it is the most outstandingly successful theory yet devised by human minds. It is almost maddeningly good: it seems
that each time a new measurement is made, it only serves to reinforce the Standard Model. The model is home of the
most accurate measurement in science (the gyromagnetic ratio of the electron). The measurement currently (B. Odom et
al., Physical Review Letters 97, 030801, 2006) has a relative error of 7.6 * 10-13, and remarkably, the theoretical
prediction is in complete agreement.
We will assume that electrons, protons and neutrons are reasonably familiar. If not, all we really need to
know before we continue is that atoms are made of them, the electron is negatively charged, the proton is
positively charged, and the neutron is neutral.
According to the Standard Model, the fundamental particles of matter are all either leptons or quarks. Both types of
particle are fermions: particles which have spin 1/2 or - 1/2 and obey the Pauli Exclusion Principle, which states that
no two can be in the same state at the same time. The theory further posits that the three fundamental forces of nature:
electromagnetism, the weak force, and the strong force (excluding gravity for the moment), are mediated by particles
called gauge bosons. The notion is that a force is felt between two charged particles when a gauge boson is exchanged
between them. They have spin +-1, so they are not fermions, and do not obey the Pauli Principle:
Lithium 7 is a boson, and Lithium 6 is a fermion. (source)

As the temperature is lowered closer and closer to absolute zero, the bosons continue to shrink together, but the
fermions are limited by the Pauli Principle.
There are two main types of leptons: electrons and neutrinos (). Except for the neutrinos, all quarks and leptons carry
electrical charge and exchange photons, which mediate the electromagnetic force. All left-handed quarks and
leptons can exchange W and Z bosons, which mediate the weak force. (A particle is left-handed when its spin is parallel
to its magnetic moment.) The weak force allows one type of quark to change into another, or an electron to change into
a neutrino (or vise versa). Only the quarks can exchange gluons, which mediate the strong force, and hold quarks
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Elementary%20Particles.html

1/6

1/3/2016

Introduction to Physical Astronomy - Elementary Particles

together. Individual quarks are never seen; they are always found as a bound pair called a meson, or a bound triplet
called a baryon. The most common baryons are protons and neutrons. The strong force is about 13.7 times stronger
than the electromagnetic force. The weak force is weaker by a factor of 10,000 than the strong force, although it is not
the weakest: gravity is 1036 times weaker. Gravity only appears to be so strong because of the sizes of the sources of
the gravitational field: planets, stars and galaxies.
There is also antimatter; for each fermion, there is an associated antiparticle. An antiparticle is identical to its ordinary
"partner" in every way except its charge and helicity (whether its spin is parallel or antiparallel to its momentum). The
antiparticle associated with the electron is the positron (e +). It is expected that issues involving handedness and helicity
will ultimately explain why there is more matter than antimatter in the observable universe.
Consider the following reaction, which lies at the heart of the nuclear reactions powering our Sun, as well as all main
sequence stars:
1H + 1H -> 2H + e+ +
e

"1H" denotes a proton: this Hydrogen nucleus has no accompanying electron to make it an atom because the intense heat
(over 10 million K) necessary to allow fusion to occur has stripped the electron off, making it an ion. "2H" denotes a
Deuterium nucleus, which consists of a proton and a neutron, bound by the strong force. The subscript on the neutrino
indicates that it is specifically an electron neutrino: the weak partner to the electron, and not to a muon or tau particle
(which are heavier leptons).
To understand this reaction, we need to know the quark content of the proton and neutron. The proton is made up of 2
"up" quarks and 1 "down" quark (up and down are just labels), bound by the strong force. The neutron is made up of
two down quarks and one up quark, similarly bound. The up quark is usually assigned an electric charge of +2/3, and
the down a charge of -1/3, of the magnitude of that of the electron. This makes the electric charge "come out even": we
call this conservation of charge.
Now that we have discussed all the pieces of the Standard Model that we will need, this diagram may help to put
everything in perspective:
Ordinary matter content of the Standard Model of Particle Physics.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Elementary%20Particles.html

2/6

1/3/2016

Introduction to Physical Astronomy - Elementary Particles

The "bar" above the second neutrino marks it as an anti-neutrino; "u" and "d" stand for up and down quarks,
respectively.
So what really happened in our fusion reaction was that one up quark turned into a down quark, via the emission of a
W+ boson (which then decayed into a positron and a neutrino):
Fusion of two protons into Deuterium.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Elementary%20Particles.html

3/6

1/3/2016

Introduction to Physical Astronomy - Elementary Particles

This allowed the resulting neutron to be bound to the remaining proton, making the Deuterium nucleus. But because
there were no leptons on the left hand side of the reaction, there must be none on the right: the positron, with lepton
number -1, plus the neutrino, with lepton number +1, add up to 0 leptons. We call this lepton number conservation.
Since protons and neutrons are both baryons, we also have baryon number conservation in this reaction.
The extra energy that is released in this reaction comes from the fact that the combined mass of the products is lower
than the combined mass of the reactants. This mass difference is equivalent to energy through Einstein's famous equation
E = m c2
and comes about because the binding energy of the resulting Deuterium nucleus is negative. Differences in binding
energies also account for the excess energy released in fission.

Theory vs. Experiment - a reality check


There is a lot of information on the Internet and on television about physics and astronomy. In many cases (even on
PBS), speculation is presented as fact. Science is at its heart very conservative, in the sense that every piece of scientific
data must be rigorously verified and replicated before it is accepted, and a theory is only accepted until it agrees with all
of the data and provides a simpler or more complete framework in which to understand that data. In a healthy scientific
environment, experiment typically drives theory:
Experiment

Theory

1814 - Fraunhofer discovers absorption spectra

1913-5 - Bohr explains them quantum


mechanically

1845 - Leverrier measures the perihelion shift of Mercury

1916 - Einstein explains using General Relativity


(GR)

1887 - Michelson and Morley determine c is a constant

1905 - Einstein explains with Special Relativity


1916 - Einstein incorporates this into GR as the

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Elementary%20Particles.html

4/6

1/3/2016

Introduction to Physical Astronomy - Elementary Particles

1893 - Eotvos determines Minertial = Mgrav

Equivalence Principle

(1800s) - measurements of black body radiation

1900 - Planck derives black body distribution

1929 - Hubble discovers the expansion of the universe

1933 Robertson finds GR solution for expanding


universe

1965 - Penzias and Wilson discover Cosmic Microwave


Background Radiation (CMBR)
1965 - Rubin begins measuring galaxy rotation curves

no consensus yet on what dark matter is

1998 - Filippenko and Riess discover the expansion of universe


no consensus yet on what dark energy is
is accelerating
When a theoretical scientist has one of those very rare, remarkable strokes of insight, and sees the existing data in a new
and fundamentally more complete way, it is possible for the conclusions of the theory to drive the search for new
experimental data:
Theory
1916 - Einstein formulates General Relativity

Experiment
1920 - Dyson, Eddington and Davidson report
gravitational deflection of light
1960 - Pound and Rebka measure gravitational red shift
1969 - Jenkins measures gravitational time effects
1978 - Taylor and Hulse measure energy loss in binary
pulsars due to gravitational waves
1980 - Hubble Space Telescope discovers gravitational
lensing

1916 - Schwarzschild discovers spherical vacuum


solution to GR (describes black holes)

no direct evidence yet

1917 - Einstein proposes Cosmological Constant

2007 - 5 years of WMAP data favors nonzero


Cosmological Constant

1921/6 - Kuluza and Klein propose extra dimensions

no evidence yet

1930 - Dirac predicts antimatter

1932 - Anderson discovers the positron

1931 - Fermi predicts the neutrino

1956 - Reines and Cowan discover it

1961 - Gell-Mann and Ne'eman explain particle zoo


using quark model

1974 - Richter and Ting discover the J/

1963 - Kerr discovers vacuum solution to GR with


angular momentum

2007 ? - Gravity Probe B searches for frame dragging

1967 - Glashow, Weinberg and Salam formulate


Electroweak theory

2013 - we seem to have a Higgs!

1972 - Gell-mann develops Quantum Chromodynamics


no contradictions yet...
(QCD)
1974 - Pati and Salam formulate first Grand Unified
Theory

no evidence yet

1975 - Hawking proposes that black holes emit


radiation

no evidence yet

1971 - Ramond, Neveu and Schwartz formulate


Supersymmetry

no evidence yet

1975/81 - Scherk, Green and Schwarz develop String

no evidence yet

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Elementary%20Particles.html

5/6

1/3/2016

Introduction to Physical Astronomy - Elementary Particles

Theory
1981-2 - Guth, Linde, Albrecht and Steinhardt propose
no contradictions yet...
inflation models
1986 - Ashtekar develops loop quantum gravity

no evidence yet

Note: Electroweak theory and QCD are incorporated into the Standard Model.
It is hoped that this little "experiment vs. theory" reality check will clarify some of the ideas presented in the public forum
in a properly scientifically conservative light.
Table of Contents
References
Index
2014, Kenneth R. Koehler. All Rights Reserved. This document may be freely reproduced provided that this
copyright notice is included.
Please send comments or suggestions to the author.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Elementary%20Particles.html

6/6

1/3/2016

Introduction to Physical Astronomy - Expansion of the Universe

Expansion of the Universe


Doppler Shifts and the Hubble Relation
(View Cosmos DVD 6, episode 10, dramatization of red shift discoveries at Mt. Wilson.)
In 1998 and 1999, the GHASP survey of spiral and irregular galaxies (source) measured the central wavelength
and width of the Hydrogen- line at 6562.78 Angstroms for the following galaxies in the UGC catalog (values
were derived from the paper for the purposes of this text):
UGC #

central (A)

diameterangular (am)

width (A)

2034

6575.29

2.43286

2.818

-17

12.9

2080

6582.36

13.1647

5.754

-19.5

11.2

3574

6594.17

6.98982

3.467

-17.6

14.1

4325

6573.78

4.62356

3.388

-17.7

12.1

4499

6577.66

3.3546

2.344

-16.2

14.1

5253

6591.66

11.3157

4.168

-20.8

10.8

5316

6585.16

5.77299

4.073

-19.7

11.5

5721

6574.37

4.2076

1.819

-16.9

12.6

5789

6578.75

5.30686

5.888

-19.2

11.2

5829

6576.52

2.10449

5.128

-17.1

12.9

5931

6597.89

8.18138

1.949

-20.1

11.7

5935

6599.49

2.66179

3.715

-20.3

11.5

5982

6597.1

9.7857

4.073

-20.3

11.5

6778

6583.74

13.0248

4.073

-20.6

10.4

7524

6569.74

3.10968

13.8

-17.3

10.6

8490

6566.96

3.83073

4.677

-17.7

11.1

9969

6617.91

14.6256

5.128

-21.7

11.2

10310

6578.2

3.70572

2.884

-17.4

13.1

12060

6582.29

4.21267

1.318

-16

14.8

12754

6579.01

5.92111

4.073

-18.9

11.5

(The angular diameter is taken to be the major axis length reported by SIMBAD.)
Let's take a closer look at UGC 2034. Using the relation for Doppler Shift:
apparent wavelength / actual wavelength = 1 + recession velocity / c
we can solve for the recession velocity
v = c * (central / 6562.78 - 1)
= c * (6575.29 / 6562.78 - 1)
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Expansion%20of%20the%20Universe.html

1/5

1/3/2016

Introduction to Physical Astronomy - Expansion of the Universe

= 572000 m / s
Using our relation between absolute and apparent magnitude, we can solve for the distance
distancepc = 10(5 + apparent magnitude - absolute magnitude) / 5
= 10(5 + 12.9 - (-17)) / 5
= 9.55 * 106 pc
If we do this for all of these galaxies, and plot the recession velocities vs the distances, we get the following plot:
Recession velocity vs. distance for the galaxies in the GHASP survey.

The red line is the best linear fit to the data points, and its slope is called the Hubble Parameter:
H0 = 69.9 km/s / Mpc
(this value is the current "best" value; see the References; the fit for our 20 galaxies had a slope of 70.783
km/s/Mpc).
It is common when dealing with distances on this scale to use the Hubble Relation
recessional velocity = H0 * distance
to compute the distance.
Portfolio Exercise: Recreate the Hubble Graph above for our sample of 20 galaxies.
It is important to realize that our observation does not imply that we are at the "center of the universe". In the
following animation, we first see the effect of the expansion of the universe on a set of galaxies (marked by yellow
dots). Then for four sample galaxies, we draw lines to all of the other galaxies, color-coded by distance (greater
distances appear more red). It is clear that from any galaxy's point of view, all of the other galaxies are receding,
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Expansion%20of%20the%20Universe.html

2/5

1/3/2016

Introduction to Physical Astronomy - Expansion of the Universe

with those most distant receding fastest:

The Hubble Parameter has a value of approximately 69.9 km/s/Mpc. This is numerically equal to 2.265 * 1018/s. So its inverse, 1/H , has a value of 4.415 * 1017 s, or 13.99 billion years. Up to a few technicalities (which
0
reduce the value to 13.78 billion years) this is the age of the universe.
(View Cosmos DVD 1, episode 1, on the Cosmic Calender.)
Dealing with cosmological distances is quite a more difficult matter. As we have seen, the universe is expanding,
and the rate of expansion increases with the distance. The expansion is described by the scale factor, denoted
"a"; if rnow is the distance between two objects now (when, by definition, a is 1), then the distance between those
objects at a different time is
rthen = rnow * athen / anow
= rnow * athen.
The scale factor is related to the red shift (denoted "z") by the equations
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Expansion%20of%20the%20Universe.html

3/5

1/3/2016

Introduction to Physical Astronomy - Expansion of the Universe

z = apparent wavelength / actual wavelength - 1


= anow / apast - 1
So we see that, as with any distance, wavelengths stretch with the expansion of the universe, and the current red
shift has a value of zero. If the Big Bang model of the universe is correct, and the universe began with infinitesimal
volume, the red shift at the Big Bang is infinite.
Comparing the equation for red shift to our equation for the Doppler Shift, we see that when the recessional
velocity is equal to the speed of light, the red shift is 1. Near this region the Hubble Relation breaks down, and
relativistic effects must be taken into account. For this reason, it is most appropriate for us to discuss distances
not in terms of meters or light years or parsecs, but instead in terms of red shift, which is based on direct
observation, and does not depend on the scale factor of the cosmological model involved.

The Accelerating Universe


This is a plot of the log of the distance (as measured from the apparent and absolute magnitudes) of highly red
shifted Type Ia supernovae, as a function of their red shift:
Log(distance) vs. red shift of high-z Type Ia supernovae. (source)

The data follows the upper line, which is a fit to a model in which the expansion of the universe accelerates due to
something called the Cosmological Constant: a constant energy density which acts like a negative pressure.
In General Relativity, pressure acts like mass or energy, and so causes gravitational attraction.
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Expansion%20of%20the%20Universe.html

4/5

1/3/2016

Introduction to Physical Astronomy - Expansion of the Universe

Negative pressure therefore causes repulsion.


This was the first evidence that the expansion of the universe is accelerating. As we shall see, further data
collected by the Wilkins Microwave Anisotropy Probe (WMAP) indicates that the Cosmological Constant may
not have been Einstein's greatest mistake after all!
Table of Contents
References
Index
2013, Kenneth R. Koehler. All Rights Reserved. This document may be freely reproduced provided that this
copyright notice is included.
Please send comments or suggestions to the author.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Expansion%20of%20the%20Universe.html

5/5

1/3/2016

Introduction to Physical Astronomy - Galaxies

Galaxies
Galaxy Morphology
Note that the classification of a given galaxy is not always universally agreed upon. The classifications below follow the morphological types quoted in SIMBAD:
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows system, download the Java VM (Virtual Machine) for your
version of Windows at the download section at java.sun.com.
Spiral Galaxies
Considering spiral types a c, the central bulge is smaller, arms are more separated and the age is greater in type c.
Most disc stars in our galaxy have completed over 40 orbits around the galactic center.
Galaxies with thick discs are typically older than those with thin discs.
Stars in the bulge date from the formation of the galaxy; star formation in the disc is ongoing. The stellar halo (about 1% of the galaxy mass) is the oldest part.
Spirals are primary sites for star formation; short-lived stars can't migrate far, so the birth rate is greater.
Near IR (1-5 m) is due mostly to red giants, so it tracks mass; blue light is mostly from young stars and so tracks star formation.
Interstellar gas is essential for persistent spiral structure, but spirals are probably not permanent; they continually dissolve and re-form. Gas can lose energy but
not angular momentum; spiral arms form since they allow transfer of angular momentum outward and mass inward.
Most spirals have two arms, and most trail along the direction of galaxy rotation, so whatever creates spirals must not be time-reversible.
Barred Spiral Galaxies
Approximately 30 to 50% of all spirals are barred.
Spirals commonly start at the ends of the bars; there is usually rapid star formation there.
Elliptical Galaxies
Ellipticals are often subclassed from E0 (nearly spherical) to E7 (very elongated); the eccentricity of an En galaxy is n/10.
Elliptical galaxies have almost no interstellar gas or dust; therefore there is no ongoing star formation. The stellar populations of most ellipticals are very old, on
the order of the age of the universe.
Lenticular Galaxies
Lenticular galaxies have type "S-zero".
Lenticular galaxies have a disc but no gas or spiral.
Irregular Galaxies
Some references (including SIMBAD) do not distinguish between Irregular types, or use distinctions other than I and II.
Irregulars are very common, and rich in gas.
They are often formed from galaxy interactions (see below).
Ring Galaxies
Ring galaxies are very rare (about 1 in 10000); they are formed by head-on collisions with a disc galaxy.
Interacting Galaxies
NGC 4013 has a tidal stream.
The Hercules Cluster indicates that galaxy interactions may be common.
Interactions and mergers probably contribute to differences in morphology.
When galaxies collide, their gasses do but their stars do not; stellar orbits will, however, change. The colliding gasses feed star formation.
This simulation of galaxy interactions illustrates how unusual morphologies might form over cosmological time scales: it spans over a billion years,
and produces a result comparable to NGC 7600. (source)
Dwarf Galaxies
Some references also consider dwarf galaxies as a separate morphological type.
Note that many galaxies (for example, Centaurus A) when viewed in multiple wavelengths challenge our notions of galactic morphology gleaned from visible light
observations.
The effective radius of a galaxy as measured by its interstellar gas is much larger than its visible radius; the low density in that region inhibits star
formation. The 21 cm Hydrogen line is observable far beyond the stellar distribution. 21 cm discs are frequently warped.
Also, see NGC 2915: a radio spiral.
(View Cosmos DVD 6, episode 10, Milky Way and galaxy interaction simulations.)
Also see this site at the University of Hawaii.
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Galaxies.html

1/4

1/3/2016

Introduction to Physical Astronomy - Galaxies

Portfolio Exercise: Find an additional example of each morphological type (Sa, Sb, Sc, SBa, SBb, SBc, E, S0, SB0, I, Ring and Interacting) and
verify the type using SIMBAD. Include an image and as much information as you can (including all URLs used) for each example.
The Sersic law is a useful model for the surface brightness (luminosity per unit area) of a galaxy:
Sersic law with re = 1 and I(re) = 1.

1/m -1)

I(r) = I(re) e-bm ((r / re)

where re is the effective radius (containing half of the total luminosity), m is the Sersic index (6 for luminous ellipticals, 2 for dim ones, 1 for disc galaxies) and bm is
an experimental fit, here taken to be 2m - 0.324.
Portfolio Exercise: Using the mass-luminosity relation, plot the mass as a function of radial distance for the parameters plotted above.
Note that this will be the total mass contained within a radius r.

Central Black Holes


Since 2001 it has been known that there is a correlation between the mass of supermassive central black holes in galaxies and the mass of the central bulge: the
median black hole mass is .0013 times the bulge mass. This correlation hints that the events which lead to the formation of supermassive central black holes are the
same as the events which lead to the formation of central bulges (in spiral or elliptical galaxies). (source)
More recently, it has been discovered that this ratio is 20 to 30 times larger for galaxies at high red shift (z > 6). (source) This seems to indicate that the black holes
form before the central bulges. Interesting problem!
Note: massive black holes formed within 10 Kpc of a galactic center will spiral in to the center within the age of the universe.
In a "cold" disc, conservation of angular momentum keeps a central black hole from accreting matter; with nothing to speed up the matter, its distance from the center
of rotation must stay constant. During mergers, however, both the black hole and the central bulge may grow.

Clusters of Galaxies
Here is a segment of the Coma Cluster (the foreground star in the upper right is HD 112887):
The Coma cluster in optical, with labels, and in xrays. (source)

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Galaxies.html

2/4

1/3/2016

Introduction to Physical Astronomy - Galaxies

(source)
Note that much of the x-rays emanate from a region which is visually devoid of matter; although this radiation comes from hot gases which are found in most clusters,
it is not dark matter.
The distribution of galaxies is homogeneous on scales 100 Mpc.
Clusters of galaxies in the early universe seem to have a larger percentage of spirals than nearer clusters.
Dense clusters have a larger percentage of ellipticals.
Lenticular galaxies are more common in the centers of clusters.
Galaxy clusters are young; most members have made very few orbits about the center of mass.
Collisions between galaxies in a cluster are much more frequent than those between stars in a stellar cluster.
Most of the baryonic matter in clusters is in the hot gasses between galaxies.
Galaxies orbit the center of mass of their clusters. In doing so, they can interact with the intracluster medium. A particularly clear example of this is ESO 137-001,
which is moving at almost seven million km/hr. In this image, the dark blue trails are x-ray data from Chandra showing the material which has been stripped from the
galaxy by ram pressure:
ESO 137-001 in optical, and in x-rays. (source)

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Galaxies.html

3/4

1/3/2016

Introduction to Physical Astronomy - Galaxies

Table of Contents
References
Index
2014, Kenneth R. Koehler. All Rights Reserved. This document may be freely reproduced provided that this copyright notice is included.
Please send comments or suggestions to the author.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Galaxies.html

4/4

1/3/2016

Introduction to Physical Astronomy - Image Processing

Image Processing
Modern astronomical images often depict colors which are far from true. But this is not just for dramatic effect: the
colors very often encode technical information about how the image was made. It is important to understand what
the colors in a given image mean, and for that reason we must discuss color filtering and compositing.
This graph shows the Johnson-Cousins photometric system. (Standard Photometric Systems, Michael S. Bessell,
Annual Review of Astronomy and Astrophysics 43 293-336 (2005))

Also denoted the ubvri photometric system, it breaks the electromagnetic spectrum into five bands which
correspond to the wavelengths passed by filters in the near ultraviolet ("u"), blue ("b"), "visual" ("v"), red ("r") and
near infrared ("i") ranges. The v band is particularly important, for it is often used to determine the apparent
magnitude. At a distance of 10 pc from a star, the apparent magnitude
m = 4.83 - 2.5 * log10 luminositysolar
where the luminosity is the total energy output in the v band, relative to that of the Sun. Often m is computed at a
specific wavelength, near 5448 A, which is an "effective" central wavelength representative of the band.
But filters on many telescopes, including the Hubble, are much more specific, concentrating on narrow emission
wavelengths corresponding to specific elements, ions and molecules. For instance, the famous image of the "Pillars
of Creation" (stellar nurseries in M16, the Eagle Nebula) was taken first as black and white exposures through
filters corresponding to:
emissions from Sulfur (S+, 6716 and 6731 A),
emissions from Hydrogen (6563 A) and
emissions from Oxygen (O++, 4363 and 5007 A).
These were then colored red, green and blue, respectively, and digitally added to produce the final image. This has
become a commonly-used color palette for images of nebulae.
The natural red color of this emission nebula arises from Hydrogen emissions at 6563 Angstroms
(called "H-Alpha"). These emissions are powered by ultraviolet radiation from hot stars, whose winds
help to create the shapes and shock waves characteristic of so many nebulae; the radiation ionizes
Hydrogen atoms, and the red emission results when the ionized electrons recombine with the
Hydrogen nuclei. Light from blue reflection nebulae, in contrast, is reflected from dust grains which,
like our atmosphere, scatter blue light more efficiently than other colors. M42 (The Orion Nebula) and
M20 (The Trifid Nebula) provide good examples of both types, as well as molecular dust lanes.
This Java applet shows how red, green and blue images can be composited back into a multicolor image. It is an
illustration of the RGB color model: by adding different amounts of the three fundamental colors, you can produce
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Image%20Processing.html

1/2

1/3/2016

Introduction to Physical Astronomy - Image Processing

(in this program) one million different shades ("True Color" on a computer monitor is considered to be 224 =
16,777,216 shades, although professional photographers work with 248, or over 281 trillion shades!).
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows system,
download the Java VM (Virtual Machine) for your version of Windows at the download section at java.sun.com.
It is also possible to enhance contrast in an image without modifying the color content. In addition to the examples
above, false color can also be used to composite images from multiple instruments:
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows system,
download the Java VM (Virtual Machine) for your version of Windows at the download section at java.sun.com.
Table of Contents
References
Index
2012, Kenneth R. Koehler. All Rights Reserved. This document may be freely reproduced provided that this
copyright notice is included.
Please send comments or suggestions to the author.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Image%20Processing.html

2/2

1/3/2016

Introduction to Physical Astronomy - Motion in the Solar System

Motion in the Solar System


This applet knows where some things in the Solar System (and beyond) are. You can use it to explore a number of astronomical
phenomena.
You can choose where you are from the first pop-up menu, and what you are looking at from the second. If the two
are not the same, you can choose the latitude where you are; if they are the same, the latitude scroll bar becomes a "z"
scroll bar, allowing you to view the body from above or below. When the "from" and "viewing" targets are different
and you are viewing from a planet, you may choose either geocentric or heliocentric views. The applet begins in
"heliocentric" mode, meaning you are looking at the sky in flat projection, with the latitude relative to the plane of the
Solar System; as you change the date and time, or animate, you will always be at the first object, viewing the second
object. In "geocentric" mode, you are viewing from the first object, but seeing the sky in polar projection (as if you
were standing on the Earth).
The date and time scroll bars allow you to see the sky at a specific date and time. You can also draw and erase lineof-sight markers using the appropriate buttons, but they are only drawn when the "from" and "viewing" targets are the
same.
When the "from" and "viewing" targets are different, the coordinates represent angles and the lattice lines are drawn
every 15 degrees. When the "from" and "viewing" targets are the same, the coordinates are distance. All times are
Universal Time.
It's a little complicated, but it's a powerful little program. Even the author hasn't fully explored its possibilities...
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows system, download the
Java VM (Virtual Machine) for your version of Windows at the download section at java.sun.com.
You can use it to explore the following phenomena:
the ecliptic
From Earth, view Alpha Centauri in Heliocentric mode and press the "Animate" button twice; you will see the
Sun, the Moon and the planets move along the center line: the plane of the Solar System, also called the ecliptic;
pressing the "animate" button one more time stops the animation.
daily, monthly and seasonal events
In Geocentric mode, at a latitude of 39 degrees, push the "Hour" button repeatedly; observe the motion of the
planets and stars as they "rise" in the East and "set" in the West.
Then, press the "Day" button repeatedly and observe the changes in the midnight sky as the days of a month
progress; in Heliocentric mode, from Earth view the Earth and press the "Magnify" button; see how the phases
of the Moon change as you repeatedly press the "Day" button.
Reload the page; in Geocentric mode, at a latitude of 39 degrees, push the "Magnify" button and set the hour to
12; then observe the changes in the position of the noonday Sun as you repeatedly press the "Month" button;
the path over the course of a year is called an analemma.
Reload the page; from the Sun, view the Sun; set the z control to 24 and push the "Animate" button twice;
observe the relative positions of the planets (through Saturn) as they orbit the Sun.
eclipses
Reload the page; in Heliocentric mode, press the "Magnify" button, then press the "Animate" button twice and
watch the Moon move in and out of the plane of the ecliptic; as the Moon moves in front of or behind the Sun,
we observe partial or full eclipses based on where the Moon lies relative to the ecliptic.
retrograde motion
From Earth, push the "Animate" button twice while viewing Betelgeuse on 6/1/00 (pay attention to Mercury) or
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Motion%20in%20the%20Solar%20System.html

1/6

1/3/2016

Introduction to Physical Astronomy - Motion in the Solar System

10/1/07 (Mars); or view Fomalhaut on 7/1/03 (Mars), or M48 on 8/1/05 (Mercury, then Mars, then Saturn).
Now view Mars, set the date for 8/1/07 and alternately push the "Draw Marker" and "Month" buttons until
5/1/08. Then from the Sun view the Sun and set "z" to 4; the markers show how the position of Mars changes
relative to the stars as Earth "laps" Mars.
solar vs. sidereal days
Noon is defined as the time when the Sun is highest in the sky (at its zenith); a solar day is the time between
two successive noons. A sidereal day is the time between two successive risings of any given star (other than
the Sun).
Reload the page; from Earth, view the Sun and push the "Draw Marker" button (nothing will appear yet); then
view Altair and draw a new marker; then press the day button once and again draw two markers as before;
finally, from the Sun view the Sun and set "z" to 3; since the two markers pointing to Altair are parallel, but the
two pointing to the Sun are not, you can see that in 24 hours, due to the change in the position of the Earth as it
rotates around the Sun, the position of the Sun changes relative to the stars; this causes the sidereal day to differ
from the solar day.
synodic vs. sidereal months
A synodic month is the time between two full moons.
Reload the page; from Earth, view the Earth and push the "Magnify" button; push the "Day" and "Hour" buttons
until the Moon is aligned along one of the axes. Then set the time for 27.3 days (1 sidereal month) later and note
the position of the Moon (the slight difference is due to the Earth's rotation around the Sun); then set the time for
2.2 days later than that (one synodic month since the first time) and note the position of the Moon, and that its
phase is the same as it was 29.5 days earlier.
These GOES 2/26/98 solar eclipse images provide an interesting perspective. (source)

Here is a time-lapse sequence of the Moon through a complete lunar cycle.

Solar System Observation


With careful telescopic observation, we can learn a number of important parameters about some of these bodies in our Solar
System:
Body

Perihelion

Aphelion

Ang. Diam. @

Orbital

Orbital

Rotational

(AU)

(AU)

ca. (as)

Period (Yrs)

Inclination

Period (Days)

Sun

1918

2.2 * 108

Axial Tilt

25.38

7.25

Mercury

0.3075

0.4667

13.03

0.241

58.65

Venus

0.718

0.728

65.44

0.615

3.39

-243

177.3

Earth

0.983

1.017

5. * 10-5

0.9973

23.45

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Motion%20in%20the%20Solar%20System.html

2/6

1/3/2016

Introduction to Physical Astronomy - Motion in the Solar System

The Moon

0.00243

0.00271

1968

0.075

Mars

1.381

1.666

25.73

1.881

10-5

Phobos

6.175 *

Deimos

1.567 * 10-4

Ceres

10-5

10-4

5.145

27.32

6.68

1.8

1.026

25.19

1.026

0.1047

8.731 *

1.569 * 10-4

0.05824

0.003456

1.8

1.026

2.549

2.9853

1.6658

4.6

10.58

0.379

Eros

1.133

1.783

0.2022

1.76

10.83

0.2202

Gaspra

1.8251

2.5929

0.0276

3.29

4.1

0.2942

Ida

2.732

2.99

0.06385

4.84

1.14

0.1936

Mathilde

1.9422

3.2082

0.08541

4.31

6.71

17.45

Vesta

2.1506

2.5734

1.2517

3.63

7.14

0.2232

Jupiter

4.952

5.455

50.1

11.86

1.305

0.4135

Io

0.00281

0.002832

1.2761

0.004844

0.04

1.769

Europa

0.004445

0.004526

1.097

0.009723

0.47

3.551

Ganymede

0.007138

0.007167

1.846

0.01959

0.21

7.155

Callisto

0.0125

0.01268

1.684

0.04569

0.51

16.69

Saturn

9.021

10.05

20.76

29.4

2.484

0.444

Mimas

0.001215

0.001265

0.06753

0.00258

1.53

0.9424

Enceladus

0.001584

0.001598

0.0851

0.003751

1.37

Tethys

0.00197

0.00197

0.1826

0.005169

1.86

1.888

Dione

0.002517

0.002528

0.1929

0.007493

0.02

2.737

Rhea

0.00352

0.003527

0.2632

0.01237

0.35

4.518

Titan

0.007929

0.008406

0.8871

0.04366

0.33

15.95

Hyperion

0.00887

0.01093

0.03062

0.0583

0.43

13

Iaepetus

0.02313

0.02448

0.2474

0.2172

14.72

79.33

Uranus

18.29

20.1

4.086

84.02

0.77

-0.7196

Miranda

8.658 * 10-4

8.705 * 10-4

0.03769

0.00387

4.2

1.413

Ariel

0.001272

0.001281

0.09246

0.0069

0.3

2.52

Titania

0.00291

0.002923

0.126

0.02384

0.14

8.706

Neptune

29.81

30.33

2.372

164.8

1.769

0.6712

Triton

0.002371

0.002371

0.1295

-0.01609

157.3

-5.877

6.363 *

3.12

26.73

(chaotic)
97.86

29.58

("ca." stands for "closest approach".)


Since we telescopically measure only angular positions in the sky, we cannot measure perihelion and aphelion directly; we require an
independent measurement of distance. This is done, for example, by radar ranging of Venus at closest approach. Thus the perihelion
and aphelion values are actually computed and not directly observed. Planetary orbits are elliptical, and orbital parameters such as
perihelion and aphelion are obtained by fitting careful measurements to elliptical orbits.
The data in the tables in this section come from NASA and NASA. There is an Excel spreadsheet available containing some of this
information. It will serve as a starting point for your efforts to duplicate the conclusions below.

Orbits
An ellipse is parametrized by a pair of lengths. In the definition of an ellipse, one is the length between two reference points (called
foci), and the other defines the size of the ellipse: for every point on the ellipse, the sum of the distances between that point and the
two foci is a constant. We can trade these lengths for two of more physical interest: the perihelion and aphelion. But we often prefer
a different pair of values: the length of the semimajor axis (denoted "a") and the eccentricity (denoted ""), which is the ratio of
the distance between the foci to the semimajor axis length ( = 0 is a circle):
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Motion%20in%20the%20Solar%20System.html

3/6

1/3/2016

Introduction to Physical Astronomy - Motion in the Solar System

In this plot, the sum of the lengths of the two red lines is a constant for every point on the ellipse.

Using the above information and the equations


perihelion = semimajor axis * (1 - eccentricity)
aphelion = semimajor axis * (1 + eccentricity)
and
radians = * degrees / 180
diameter = distance * angular diameterradians
we can compute the following parameters:
Body

a(AU)

Eccentricity

109

R(m)
108

R(Earth)

Sun

1.62 *

Mercury

0.3871

0.2056

2.44 * 106

0.3825

Venus

0.7233

0.0068

6.052 * 106

0.9488

Earth

0.01671

6.378 * 106

The Moon

0.00257

0.0549

1.734 * 106

0.2719

Mars

1.524

0.0934

3.397 * 106

0.5326

Phobos

6.269 * 10-5

0.015

1.3816 * 104

0.002166

Deimos

1.568 * 10-4

5. * 10-4

7688

0.001205

Ceres

2.767

0.0789

4.73 * 105

0.0742

Eros

1.458

0.2229

11001

0.001725

Gaspra

2.209

0.1738

8428

0.001321

Ida

2.861

0.0451

2.025 * 104

0.003175

Mathilde

2.646

0.266

3.265 * 104

0.005118

Vesta

2.362

0.0895

2.65 * 105

0.04155

Jupiter

5.203

0.04839

7.149 * 107

11.21

Io

0.002821

0.004

1.821 * 106

0.2855

Europa

0.004485

0.009

1.565 * 106

0.2454

Ganymede

0.007153

0.002

2.634 * 106

0.413

Callisto

0.01259

0.007

2.403 * 106

0.3768

Saturn

9.537

0.05415

6.027 * 107

9.449

Mimas

0.00124

0.0202

1.96 * 105

0.03073

Enceladus

0.001591

0.00452

2.47 * 105

0.03873

Tethys

0.00197

5.3 * 105

0.0831

6.955 *

109

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Motion%20in%20the%20Solar%20System.html

4/6

1/3/2016

Introduction to Physical Astronomy - Motion in the Solar System

Dione

0.002523

0.00223

5.6 * 105

0.0878

Rhea

0.003523

0.001

7.64 * 105

0.1198

Titan

0.008167

0.02919

2.575 * 106

0.4037

Hyperion

0.0099

0.104

8.878 * 104

0.01392

Iaepetus

0.02381

0.02828

7.18 * 105

0.1126

Uranus

19.19

0.04717

2.559 * 107

4.012

Miranda

8.681 * 10-4

0.0027

2.36 * 105

0.037

Ariel

0.001276

0.0034

5.79 * 105

0.09078

Titania

0.002916

0.0022

7.889 * 105

0.1237

Neptune

30.07

0.00859

2.476 * 107

3.883

Triton

0.002371

1.6 * 10-5

1.352 * 106

0.212

Tempel 1

3103

4.865 * 10-4

Wild 2

2750

4.312 * 10-4

Of course, a and are sufficient to describe any one of these orbits, but in order to understand how the planets are oriented with
respect to each other we need 4 additional values:
the orientation of the ellipse with respect to, for instance, the direction of the galactic center;
the angle of inclination of the plane of the ellipse with respect to the plane of the Earth's orbit;
the angle of the twist of the plane of the ellipse (around the semimajor axis), and
the position of the planet on its ellipse at a specific time.
Together these 6 values are called orbital elements. From them we can predict the past and future positions of any of the planets,
to reasonable accuracy, within about 20 years of the time mentioned above. Beyond those dates, gravitational interactions between
the planets must be taken into account.
There are subtle variations in eccentricity and tilt which cause long term cycles in the amount of sunlight received in the northern
hemisphere, and which drive cyclic climate change:
These cycles were discovered by Milutin Milankovitch, a Serbian mathematician, and account for the ice ages which occur in
100000 and 41000 year cycles, as well as the smaller variations that occur in 19000 to 23000 year cycles. (source)

Orbital cycles do not account for the increases in average global temperature since 1880. (source)

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Motion%20in%20the%20Solar%20System.html

5/6

1/3/2016

Introduction to Physical Astronomy - Motion in the Solar System

Portfolio Exercise: Compute the following characteristics for each body in the table above: volume (in units of Earth's
volume), surface area (in units of Earth's surface area) and average orbital velocity (in km/hour). Include details on the
formulas used to compute the radius of each body.
Table of Contents
References
Index
2011, Kenneth R. Koehler. All Rights Reserved. This document may be freely reproduced provided that this copyright notice is
included.
Please send comments or suggestions to the author.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Motion%20in%20the%20Solar%20System.html

6/6

1/3/2016

Introduction to Physical Astronomy - Nuclear Reactions

Nuclear Reactions
The primary source of the Sun's power output is the proton-proton chain:
1H + 1H -> 2H + e+ +
e
2H + 1H -> 3He +
3He + 3He -> 4He + 1H + 1H

( denotes a photon.) The first two of these reactions occurs twice for each occurrence of the
third reaction.
In these nuclear reactions, the difference between the masses of the products and reactants can be used to
compute the energy released by using
energy = mass * c2
Most people think of this as "converting matter into energy". Since energy is a
quality of matter, and not really a separate entity, it is better to think of this process
in terms of the constituent quarks of the protons and neutrons involved. Then we
can see that each reactant has an energy content associated with the fact that they
are bound states of particles which are electrically as well as strongly interacting.
The product has a different binding energy, and the fusion process liberates the
difference as gamma rays (with electromagnetic energy) and thermal energy of
motion. The energy of any neutrinos which are produced is usually included as part
of the energy released, because the neutrino has a very small and not very
accurately measured mass in comparison with the other particles.
The mass of a proton (1H) is 1.673 * 10-27 kg; the mass of a Helium nucleus (4He) is 6.647 * 10-27 kg, and
the mass of a positron (e+) is 9.109 * 10-31 kg. The energy released in neutrinos (e) and gamma rays () is
then
E = (4 * 1.673 * 10-27 - 6.647 * 10-27 - 2 * 9.109 * 10-31) * (3 * 108)2
= 3.881 * 10-12 J
(= 24.225 MeV)
Using the solar constant (1365 W/m2) as the energy flux and 1 AU as the distance in the equation below, we
can compute the luminosity of the Sun, obtaining LS = 3.84 * 1026 W. Dividing the luminosity by the energy
per reaction gives the number of reactions per second, 9.92 * 1037. Since each reaction consumes 4 protons,
the rate of Hydrogen consumption is
9.92 * 1037 * 4 * 1.673 * 10-27
= 6.64 * 1011 kg / s.
The photons released in these reactions are gamma rays, but the bulk of the electromagnetic radiation leaving
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Nuclear%20Reactions.html

1/4

1/3/2016

Introduction to Physical Astronomy - Nuclear Reactions

the surface of the Sun is in the visible spectrum. The average travel distance (mean free path) between
collisions for a photon in the center of the Sun is about 1 cm. On average, each photon leaving the surface has
interacted with matter about 1021 times since being emitted in the core. Since many of these interactions
involve the absorption and re-emission of the photon, the photons emitted at the surface are really distant
ancestors of the ones created during fusion. Most interactions involve some energy loss (they are "inelastic"),
and this accounts for the shift in spectrum from gamma to visible. The average time for a photon to travel to
the surface from the core is
1021 * .01 m / c = 33 billion seconds
or over a thousand years.
In the following list of nuclear reactions occurring in stars, numbers in parentheses are approximate energy
yields in MeV; negative values indicate additional energy is required to make these reactions occur. Note that
these are not all of the possible reactions; any reaction which conserves energy, momentum, angular
momentum, electric charge, baryon number and lepton number can occur. Each reaction occurs with different
frequency based on how much energy is required to overcome the electric repulsion between nuclei before
the strong force takes over.
Hydrogen fusion ignites at 107 K (pp chain)
1H + 1H -> 2H + e+ + (0.38)
e
2H + 1H -> 3He + (5.5)
3He + 3He -> 4He + 1H + 1H (12.9)

This process takes place not only in the interior of main sequence stars, but
on the surfaces of white dwarfs that have accreted Hydrogen from a binary
partner. This is called a nova, and the process usually repeats over time.
or (CNO Cycle, dominant beyond 16 MK)
12C

+ 1H -> 13N + (2)


13N -> 13C + e+ + (1.1)*
e
13C

+ 1H -> 14N + (11.2)


14N + 1H -> 15O + (3.6)
15O -> 15N + e+ + (1.8)*
e
15N

+ 1H -> 12C + 4He (4.9)

Helium fusion ignites at 108 K (Triple Alpha Process)


4He + 4He -> 8Be + (-0.09)
4He + 8Be ->12C

+ (7.4)

Carbon fusion ignites at 2-6 * 108 K


12C

+ 4He -> 16O + (7.2)


12C + 12C -> 20Ne + 4He (4.7)
12C + 12C -> 23Na + 1H (2.2)
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Nuclear%20Reactions.html

2/4

1/3/2016

Introduction to Physical Astronomy - Nuclear Reactions

12C

+ 12C -> 23Mg + 1n (-2.6)


12C + 12C -> 24Mg + (14)
Again, this process does not just occur during the horizontal branch of a massive star. If a
Carbon white dwarf accretes enough material from a binary companion to raise its
internal temperature sufficiently, the white dwarf will begin Carbon fusion throughout its
mass and explode in a Type Ia supernova (Type I spectra lack Hydrogen lines; Type Ia
typically have Silicon lines).
The second to last reaction in this set is included to illustrate reactions which release
neutrons. These free neutrons will enter into neutron capture reactions below.
Oxygen fusion ignites at 1.4-2 * 109 K
16O

+ 4He -> 20Ne + (4.8)


16O + 16O -> 28Si + 4He (0.47)
Silicon fusion ignites at 3.5 * 109 K
28Si + 4He -> 32S

+ (16)
32S + 4He -> 36Ar + (6.7)
36Ar + 4He -> 40Ca + (7)
40Ca + 4He -> 44Ti + (5.1)
44Ti + 4He -> 48Cr + (7.7)
48Cr + 4He -> 52Fe + (7.9)
52Fe + 4He -> 56Ni + (8)
28Si + 28Si -> 56Ni + (29.2)
56Ni -> 56Co + e+ + (1.1)*
e
56Co

-> 56Fe + e+ + e (3.5)*

Note that baryon number conservation accounts for the drop in the number of nuclei available to fuse. For
instance, it takes 56 1H to (eventually) make just one 56Fe.
The following masses (in atomic mass units (amu); 1 amu = 1.66055 * 10-27 kg) will enable
you to compute more accurate energy yields from these reactions.
n

1.0087

14N

13.9954

32S

31.9633

1H

1.00725

15N

14.9963

36Ar

35.9576

2H

2.01355

15O

14.9987

40Ca

39.9516

3He

3.0149

16O

15.9905

44Ti

43.9476

4He

4.0015

20Ne

19.9869

48Cr

47.9408

8Be

8.00311

23Na

22.9838

52Fe

51.9338

12C

11.9967

23Mg

22.9875

56Fe

55.9206

13C

13.0001

24Mg

23.9784

56Co

55.925

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Nuclear%20Reactions.html

3/4

1/3/2016

Introduction to Physical Astronomy - Nuclear Reactions

13N

13.0019

28Si

27.979

56Ni

55.9267

You can find the atomic numbers of these atoms from the Periodic Table of the Elements.
Neutron capture can form atoms with atomic masses larger than Iron, for instance in the series of processes:
56Fe + 1n -> 57Fe
57Fe + 1n -> 58Fe
58Fe + 1n -> 59Fe
59Fe -> 59Co

+ e- + e

59Co

+ 1n -> 60Co

60Co

-> 60Ni + e- + e

But nuclear fusion ends with Iron production. This is because 56Fe fusion is an endothermic process: it takes
more energy than it releases. For instance,
56Fe + 56Fe -> 112Te

requires over 44 MeV to take place. While we have seen several such reactions above, they have proceeded
because excess energy is available in the active core of the star. Once an Iron core forms, there are no
reactions available to provide the additional energy necessary for Iron fusion to occur.
It is often said that 56Fe is the most stable nucleus. A nucleus is said to be stable if its energy per
baryon is a minimum for its atomic number. Unstable nuclei undergo radioactive decay, emitting
Helium nuclei (alpha decay, which heavier nuclei often do because they lose energy quicker),
electrons or positrons (beta decay, accompanied by neutrinos), or photons (gamma decay, if
nothing else is allowed). The reactions marked with an asterisk above are all beta decays.
Fusion stops with Iron because 56Fe has the lowest energy per baryon of any nucleus (-8.8
MeV / baryon).
Portfolio Exercise: For each of the fusion reactions above, compute the energy yield.
Table of Contents
References
Index
2010, Kenneth R. Koehler. All Rights Reserved. This document may be freely reproduced provided that
this copyright notice is included.
Please send comments or suggestions to the author.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Nuclear%20Reactions.html

4/4

1/3/2016

Introduction to Physical Astronomy - Postscript: Extraterrestrial Life

Postscript: Extraterrestrial Life


Discussions of life on other worlds usually include Drake's Equation:
number of technological civilizations in the galaxy =
rate of star formation * fraction of those with planets * number of habitable planets *
fraction with any life * fraction of those with intelligent life * fraction of those with technology *
lifetime of a technological civilization
As you can see, computing the number of technological civilizations in the galaxy is largely an exercise in guesswork, Most sources choose values for the various
factors so that, in effect, the equation becomes
number of technological civilizations in the galaxy ~ lifetime of a technological civilization
Let us ask a more tractable question: how many planets are there in the galaxy which harbor any life at all? We will need to take into account the following:
Planets too close to the galactic core will receive too much radiation and cosmic ray flux for life to develop. If the central bulge has a radius of 3 kpc, it
seems reasonable to expect that life-bearing worlds would be at least twice that far from the core.
The stars at the edge of the galaxy are older and do not contain enough of the heavier elements required for life; there simply haven't been enough
generations of stars. We will assume that the outer 3 kpc of the disc is too old.
That leaves an annulus (a disc with a circular hole in the center) of inner radius 6 kpc and outer radius 12 kpc, or roughly
(122 - 62) / ( 152)
= 48 %
of the galactic disc which may contain stars with habitable planets. Since the central bulge contains about 30% of the luminous matter in the galaxy, this
means that only about 1/3 of the stars in the galaxy may have habitable planets.
In their paper "Do extragalactic cosmic rays induce cycles in fossil diversity?", Medvedev and Melott model the variation of cosmic ray flux as the solar
system moves in and out of the galactic plane, and relate it to variations in biodiversity on Earth:
Cosmic ray flux and biodiversity on Earth. (source)

This indicates that planets lying too far out of the plane of the galactic disc may also receive too much radiation and cosmic ray flux for life to develop.
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Postscript_%20Extraterrestrial%20Life.html

1/3

1/3/2016

Introduction to Physical Astronomy - Postscript: Extraterrestrial Life

The Sun is currently about 10 pc north of the galactic plane, and moves as far as 70 pc above or below it with time. Assuming that the central half of this
variation is compatible with life (70 pc), and that the thickness of the disc is about 150 pc, this reduces our third of eligible stars to about 16%.
Planets around binary systems should have orbits too unstable for life to develop. Since about 75% of all stars are members of a binary system, we are
down to about 4% of the stars in the galaxy which we would expect to have habitable planets.
Life arose on the Earth about 1 billion years after the Sun entered the main sequence. Any planet with life should probably orbit a star which has been on
the main sequence at least 1 billion years.
Liquid water appears to be necessary for life. All life that we know is characterized by the following statement: life is the maintainence of concentration
gradients across membranes. Water is the medium in which the chemicals necessary for life are dissolved; the membranes are the walls of cells, and the
concentrations of those chemicals on either side of the cell walls is different. When the concentrations become the same and stay that way, the cell is in
static equilibrium with its environment, and activities associated with living terminate. There are anhydrobiotic organisms (which can survive without any
water in their bodies), but in the dessicated state, they simply lie inert until they are re-hydrated.
So in order for a planet to develop life, it is necessary for liquid water to exist there. We can use the following relationships to estimate the "water zone"
around a star of a given mass. From the relation
stellar lifetimeyears = 1010 / masssolar2.5
and our last consideration, the star's mass has an upper limit of about 101/2.5 = 2.5 Msolar. Since less than 1% of stars have a mass in excess of 2.5
Msolar, we can safely ignore this limit here.
Because luminosity is approximately equal to mass3.5 (in solar units), the energy flux received by the planet will be
energy flux = luminosity / (4 distance2)
= Msolar3.5 Lsolar / (4 distance2)
Assuming that the planet has surface water, its temperature must be between the freezing and boiling points of water: 273 and 373 K. Stefan's Law tells
us that in equilibrium (another assumption, but reasonable), the energy flux should be equal to the
power emitted per unit area = * temperature4
so the flux received from its star should be between
2734 = 315 W / m2
and
3734 = 1097 W / m2
Our Earth obviously violates this bound, but we have not taken into account atmospheric effects. Since the average surface temperature of the Earth is
288 K, and we receive about 1365 W / m2 from the Sun, we can estimate that the presence of an atmosphere alters these numbers by the factor
1365 / ( * 2884) = 3.5
to give us 1100 and 3840. Setting our expression above for the energy flux equal to these two values, we find that the orbital distance (semimajor axis
length) should be between
(Msolar3.5 Lsolar / (4 1100))1/2
= Msolar1.75 * 1.67 * 1011 m = 1.12 AU * Msolar1.75
and
(Msolar3.5 Lsolar / (4 3840))1/2
= Msolar1.75 * 8.93 * 1010 m = 0.6 AU * Msolar1.75
If we consider our Solar System to extend out to the orbit of Neptune, at 30 AU, and assume that is a typical size, this implies that habitable planets
would only be found in an annulus representing about
(1.122 - 0.62) / ( 302)
= 0.1 %
of the Solar System's volume. Since Earth is one of eight planets in the Solar System, we might reasonably guess that about one in a hundred stars has a
habitable planet in orbit. Since there are about 100 billion stars in our galaxy, and we have estimated that 4% of them might be eligible to have a habitable
planet, but only 1% of those do, we conclude that in our galaxy alone, there are about 40 million worlds with life of some kind.
Since there are approximately 100 billion galaxies in the universe, and about 3/4 of them are spirals, we might expect there to be as many as 3 quintillion
planets with life in the universe. Does this number have any basis in fact? We have made a number of educated guesses in this analysis, but it seems fairly
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Postscript_%20Extraterrestrial%20Life.html

2/3

1/3/2016

Introduction to Physical Astronomy - Postscript: Extraterrestrial Life

obvious that, depending on your definitions, the universe is teeming with extraterrestrial life.
This study indicates that 20% of the stars in our galaxy have habitable planets. If that is true, our analysis is low by a factor of 5.
Table of Contents
References
Index
2013, Kenneth R. Koehler. All Rights Reserved. This document may be freely reproduced provided that this copyright notice is included.
Please send comments or suggestions to the author.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Postscript_%20Extraterrestrial%20Life.html

3/3

1/3/2016

Introduction to Physical Astronomy - Preface

Preface
This is a work in progress. Of course, due to the nature of astronomy in the 21st century, it will probably
always be a work in progress...
It started as notes for an introductory course, but it soon became obvious that my students wanted to look at
pretty pictures, and I wanted to use physics to teach them how we understand the astronomical observations
we make. I'm not sure at what point I decided to turn my notes into this online text. They are a lot of work,
and after writing four previous online texts, I swore there would not be another.
So here we are. I make changes and additions to the text on a regular basis, and anticipate doing so
constantly during the 2012-2013 academic year, when I will be using the text in a course for the first time.
For those of you contemplating that course, here is your chance to have a serious impact on the text: your
questions and comments will in large part help this work in progress to reach a mature state. But I assure you,
the text will continue to grow so long as we learn more about our universe.
One important thing that I have added is a portal to the text which allows you to look at the pretty pictures,
and when you want to learn more, you can enter the text by simply tapping the "Enter" key. Try it!
The portal also appears embedded in many of the pages of the text. To navigate it, use the up
and down arrows.
You will need the Java Runtime Environment (JRE) for the portal, as well as for other applets in
the text; if you do not have it, you can download it free.
The text contains a number of parenthetical exhortations to view various video excerpts, many from Carl
Sagan's Cosmos. Cosmos is still an excellent and entertaining introduction to the universe, and I highly
recommend it.
Also, I have included a number of "portfolio exercises", which are explorations you can do to extend your
understanding of the text.
The end of each page has links to the table of contents, references, and the index. There are some links to
sites outside of the text scattered throughout it, but most of the external links are on the references page. They
are intended to be suggestions about where to go from here to learn more (and find more pictures: I can
spend hours contemplating the infinite variety of nebulae). I particularly recommend NASA's Remote Sensing
Tutorial: it is a great place for more in-depth discussions of much of what is touched on in this introductory
text. The references page also contains some links and suggestions which might aid your own observations, a
summary of the equations and constants used in the text, and a brief glossary.
Throughout the text, sources and copyright information for each image are linked as "(source)".
For the images displayed by the portal, the original source (and copyright information) can be
accessed for any image by pressing the "Home" key on your keyboard.
I never cease to be amazed at the difficulty of finding mistakes in works of this size. They seem to creep in
through the cracks between the keys on the keyboard, and once in place, my tendency to read what I expect
and not what I see makes them all but invisible. If you find any, please let me know.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Preface.html

1/2

1/3/2016

Introduction to Physical Astronomy - Preface

Astronomy is one of the best applications of physics that exists, reason enough for a text of this nature. But
beyond that, astronomy is perhaps the ultimate human pursuit which distinguishes us from our evolutionary
ancestors. In it, we cast our wonder toward infinity, and attempt to reason out the structure of existence. It
wildly and ceaselessly feeds our fascination and our imagination.
Learn and enjoy!
Table of Contents
References
Index
2012, Kenneth R. Koehler. All Rights Reserved. This document may be freely reproduced provided that
this copyright notice is included.
Please send comments or suggestions to the author.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Preface.html

2/2

1/3/2016

Introduction to Physical Astronomy - Solar System Dynamics

Solar System Dynamics


Using Newton's version of Kepler's third law:
orbital period = 2 * * (semimajor axis3 / (G * mass))1/2
we can compute the mass of any body with an orbiting satellite using the satellite's orbital parameters. But since we are able to send probe
satellites to these bodies, we are able to measure their masses more accurately by measuring their gravitational influence on the paths of our
probes and using Newton's Equation:
force due to gravity = G * mass1 * mass2 / distance2
It is enlightening to see how Newton's Equation implies Kepler's laws. Bear with us:
We would like to see how Newton provided a simple and elegant mathematical framework encompassing
Kepler's Laws. So we will apply Newton's Law of Gravity to a simple system of a planet and a star, and see if
we can derive Kepler's Laws from them.
Let us first assume that the mass m is orbiting in the gravitational field created by the mass M, and that m is
spherical and far enough from M that it can be considered a point mass. Considering the enormous distances
between them relative to their sizes, this is not a problem.
Newton's law of gravity is written
F = - G m M / r2
or, using his Second law (force = mass * acceleration), as
m a = - G m M / r2
From this we see immediately that Newton's law of gravity DOES NOT DEPEND ON m .
In Einstein's General Relativity, this is a result of the Equivalence Principle, which states that Gravity acts the same
on all masses regardless of their composition: the mass's acceleration does not depend on the mass itself.
Rearranging, we then have
(G M) / r2 + a = 0
But there is another piece of information that is hidden in Newton's law of gravity: THE FORCE ACTS ON A
LINE BETWEEN THE TWO MASSES.
Since two points define a plane, this is a planar problem, meaning that we have to think in terms of x and y
directions. We lose no generality by choosing a coordinate system so that M is at the origin and m is at {x[t],
y[t]}. Since acceleration is the rate of change of the rate of change of the position with time, we write a as
{x''[t], y''[t]}
(where each prime indicates "rate of change") and multiply the (G M)/r2 term by the direction {x[t], y[t]}/r, where
r is (x2+y2) (using the Pythagorean theorem). Newton's law then becomes a pair of equations:
G M x[t] / (x[t]2 + y[t]2)3 + x''[t]=0,
G M y[t] / (x[t]2 + y[t]2]3 + y''[t]=0
This is a miserable pair of equations to solve, but if we express them in terms of r and the angle between the line
connecting m and M, and the x axis:

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Solar%20System%20Dynamics.html

1/7

1/3/2016

Introduction to Physical Astronomy - Solar System Dynamics

we can attack them more simply.


First recall that x/r = cos and y/r = sin . Making these substitutions into our pair of equations and using the
calculus that Newton invented for the occasion gives us two equivalent equations that are much more tractable:
G M / r[t]2 - r[t] '[t]2 + r''[t]=0,
r[t]2 '[t] = J
In the second equation, J is a constant called the angular momentum. It measures both the rate of rotation and
the bulk of the rotating object. Since the area of a triangle is one half the base times the height, r[t]2 '[t] is twice
the rate at which area is swept out by the path r[t]:

So we have just found Kepler's second law as a result of Newton's law of gravity: the area swept out in any
given time is a constant.
With some significant but perfectly legal manipulations (involving more of Newton's calculus), we can make the
substitution u=1/r, trade t dependence for dependence, and change the other equation to
u''[] + u[] = G M / J2
This has the solution (after we substitute 1/r for u)
r = ((1 + ) r0) / (1 + Cos[])
Here, we have traded the two physical constants describing the motion (M and J) for two geometrical constants
which describe the orbit: (G M)/J2 is 1/(r0(1+)), where r0 is the perihelion and is the eccentricity of the ellipse
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Solar%20System%20Dynamics.html

2/7

1/3/2016

Introduction to Physical Astronomy - Solar System Dynamics

describing the orbit.


As expected, M is at one focus: the origin. So we have found Kepler's first law as well.
Since the rate at which area is swept out is J/2, the time it takes to complete one orbit (the period) is the area of
the orbit ( a b) divided by J/2. Now the length of the semimajor axis (a) is (r[0]+r[])/2, or
(r0 + r0 (1 + ) / (1 - )) / 2 = r0 / (1 - )
and the length of the semiminor axis (b) is a (1 - 2). The period of the orbit is then
(2 a b) / J = (2 r02 (1 - 2)) / (J (1 - )2)
Since J is (G M r0 (1 + )), the period is
2 (1 / (G M)) (r0 / (1 - ))(3/2) = 2 (a3 / (G M))
and we have just found Kepler's third law: the orbital period is proportional to the square root of the cube of the
semimajor axis.
There is something else pretty special here. It turns out (Bertrand proved it) that the inverse square force law is
one of only TWO force laws which allow closed orbits. And the inverse square law is a direct consequence of the
fact that we live in 3 dimensions.
Now this was all for one planet and one star, but Newton's laws are powerful enough for us to use them for more
complex systems; just not in this text.
And it turns out that we have actually accomplished more than we thought we did. All this time you thought (admit
it!) that M was the mass of the star. But since the gravitational force is felt by both the star and the planet, the
planet will cause the star to wobble, so it cannot be fixed at any point. (This is how we most often detect
extrasolar planets.)
What we have actually done is find both the motion of the planet and of the star: each orbits in an ellipse around
their common center of mass, and M is the total mass (star plus planet). The origin is the location of the center
of mass of the star-planet system. Since the star is so much more massive than the planet, the center of mass is
very close to the center of the star, so we observe the elliptical orbit of the star as the wobble we spoke of above.
What about our treatment of the planet as a point mass? Gauss' law tells us that the gravitational force due to a
spherically symmetric mass is equivalent to the force due to a point particle of the same mass, located at the
center. Therefore we will usually treat planets (and stars) as if all their mass was located at a point at their center.
Using the masses of these bodies, we can compute a number of interesting parameters: their average densities, their surface gravities (by
setting "mass2" to 1 and "distance" to the radius), their orbital angular momenta ("L", using the equation
angular momentum = 2 * * mass * rotation rate * orbital radius2)
and their central pressures. The last we can estimate from dimensional considerations, using the formula
pressure = G * mass2 / radius4
Note that this value of the Earth's central pressure is 14.23 million times the atmospheric pressure on the surface.
Body

M(kg)

M(Earth)

Density

g(Earth)

L (kgm2/s)

L(Earth)

(g/cm3)

Central Pres.

Central Pres.

(Pa)

(Earth)

Sun

1.989 * 1030 3.33 * 105

1.411

28

1.424 * 1056

5.349 * 1015

1.128 * 1015

784.1

Mercury

3.302 * 1023 0.05528

5.429

0.3778

9.149 * 1038

0.03437

2.054 * 1011

0.1427

Venus

4.868 * 1024 0.815

5.244

0.9053

1.846 * 1040

0.6934

1.179 * 1012

0.8195

Earth

5.974 * 1024 1

5.496

2.662 * 1040

1.439 * 1012

The Moon 7.348 * 1022 0.0123

3.365

0.1664

2.884 * 1034

1.083 * 10-6

3.986 * 1010

0.0277

6.418 * 1023 0.1074

3.909

0.3788

3.531 * 1039

0.1326

2.064 * 1011

0.1435

Mars

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Solar%20System%20Dynamics.html

3/7

1/3/2016

Introduction to Physical Astronomy - Solar System Dynamics

Phobos

1.063 * 1016 1.779 * 10-9

3.792 * 10-4 2.132 * 1026

8.009 * 10-15

2.07 * 105

1.438 * 10-7

Deimos

2.38 * 1015

3.984 * 10-10 1.25

2.742 * 10-4 7.545 * 1025

2.835 * 10-15

1.082 * 105

7.52 * 10-8

Ceres

8.7 * 1020

1.456 * 10-4

1.963

0.0265

2.424 * 10-4

1.009 * 109

7.013 * 10-4

Eros

6.69 * 1015

1.12 * 10-9

1.2

3.765 * 10-4 3.6 * 1031

1.353 * 10-9

2.039 * 105

1.417 * 10-7

Gaspra

1016

1.674 * 10-9

3.987

9.587 * 10-4 6.609 * 1031

2.483 * 10-9

1.323 * 106

9.192 * 10-7

Ida

1017

1.674 * 10-8

2.875

1.661 * 10-3 7.536 * 1032

2.831 * 10-8

3.969 * 106

2.758 * 10-6

Mathilde

1.033 * 1017 1.729 * 10-8

0.7089

6.599 * 10-4 7.083 * 1032

2.661 * 10-8

6.266 * 105

4.355 * 10-7

Vesta

3 * 1020

3.849

0.0291

2.055 * 1036

7.719 * 10-5

1.218 * 109

8.464 * 10-4

Jupiter

1.898 * 1027 317.8

1.24

2.53

1.932 * 1043

725.8

9.209 * 1012

6.4

Io

8.932 * 1022 0.01495

3.531

0.1834

6.538 * 1035

2.456 * 10-5

4.841 * 1010

0.03364

Europa

4.8 * 1022

2.989

0.1335

4.426 * 1035

1.663 * 10-5

2.563 * 1010

0.01781

Ganymede 1.482 * 1023 0.02481

1.936

0.1455

1.725 * 1036

6.479 * 10-5

3.044 * 1010

0.02116

Callisto

1.076 * 1023 0.01801

1.851

0.1269

1.662 * 1036

6.245 * 10-5

2.317 * 1010

0.0161

Saturn

5.685 * 1026 95.17

0.62

1.066

7.837 * 1042

294.4

1.635 * 1012

1.136

Mimas

3.75 * 1019

6.278 * 10-6

1.189

0.006648

9.96 * 1031

3.742 * 10-9

6.359 * 107

4.419 * 10-5

Enceladus

7. * 1019

1.172 * 10-5

1.109

0.007814

2.105 * 1032

7.907 * 10-9

8.785 * 107

6.105 * 10-5

Tethys

6.27 * 1020

1.05 * 10-4

1.005

0.0152

2.097 * 1033

7.879 * 10-8

3.325 * 108

2.311 * 10-4

Dione

1.1 * 1021

1.841 * 10-4

1.495

0.02389

4.163 * 1033

1.564 * 10-7

8.21 * 108

5.706 * 10-4

Rhea

2.31 * 1021

3.867 * 10-4

1.237

0.02695

1.033 * 1034

3.881 * 10-7

1.045 * 109

7.263 * 10-4

Titan

1.346 * 1023 0.02252

1.881

0.1382

9.161 * 1035

3.442 * 10-5

2.748 * 1010

0.0191

Hyperion

8 * 1017

1.339 * 10-7

0.2729

6.912 * 10-4 5.993 * 1030

2.251 * 10-10

6.875 * 105

4.778 * 10-7

Iaepetus

1.6 * 1021

2.678 * 10-4

1.032

0.02114

1.86 * 1034

6.989 * 10-7

6.428 * 108

4.467 * 10-4

Uranus

8.685 * 1025 14.54

1.237

0.9032

1.696 * 1042

63.73

1.174 * 1012

0.8157

Miranda

6.6 * 1019

1.105 * 10-5

1.199

0.00807

5.728 * 1031

2.152 * 10-9

9.371 * 107

6.512 * 10-5

Ariel

1.35 * 1021

2.26 * 10-4

1.66

0.02742

1.42 * 1033

5.336 * 10-8

1.082 * 109

7.52 * 10-4

Titania

3.53 * 1021

5.909 * 10-4

1.716

0.03863

5.613 * 1033

2.109 * 10-7

2.147 * 109

0.001492

Neptune

1.024 * 1026 17.15

1.61

1.138

2.505 * 1042

94.09

1.862 * 1012

1.294

Triton

2.14 * 1022

2.067

0.07973

-3.333 * 1034

-1.252 * 10-6

9.146 * 109

0.006356

5.022 * 10-5

0.008035

0.003582

0.9623

6.453 * 1036

Several densities of interest:


liquid water at 4 C - 1 g / cm3
gaseous H2 at 0 C and 1 atm - .00009 g / cm3
metallic H - 1.3 to 1.6 g / cm3
silicate minerals - 3 to 4 g / cm3
ferric/ferrous minerals - 7 to 8 g / cm3
Two other quantities worth noting are the orbital velocity (obtained from equating the gravitational acceleration to the acceleration v2/r
required to keep a body moving in a circular orbit):
orbital velocity = (G * mass / distance)1/2
and the escape velocity (obtained from equating the kinetic energy mv2/2 to the gravitational energy Gm/r):
escape velocity = (2 * G * mass / distance)1/2
In both of these equations, the mass is the mass of the body we are orbiting or escaping from; the results are independent of the mass of our
spacecraft. The distance in each equation is the distance from the center of the body (its radius in the case of the escape velocity).

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Solar%20System%20Dynamics.html

4/7

1/3/2016

Introduction to Physical Astronomy - Solar System Dynamics

Portfolio Exercise: Use the relation between the orbital period, the semimajor axis length and the total mass, to compute the
masses of the Sun and those planets with moons.
Recall from our derivation of Kepler's Laws from Newton's that the mass in this relation is the combined mass of
the parent and the satellite. For this reason, you will get the best results if you choose the satellite with the lowest
mass (ie., use Mercury's orbital parameters to compute the mass of the Sun, Deimos' to compute Mars' mass,
etc.).
Compare these values to the measured values quoted above: in terms of a percentage, how far off are your values from those
in the table above?
For each body in the table above, compute the orbital velocity at a distance of 400 km from the surface, and the escape
velocity from the surface. Compute each value in both km/hr and in multiples of the Earth's value.

Gravity and Tides


Using the masses above, it is interesting to compare the strengths of the gravitational forces among some of these bodies. This table gives the
gravitational force between some of the planets and their moons, expressed as a fraction of the gravitational force between the Sun and that
planet:
Planet

Fgrav (Sun)

Body

Earth

The Moon

0.005594

Jupiter

Io

0.1528

Jupiter

Europa

0.03248

Jupiter

Ganymede

0.03943

Jupiter

Callisto

0.009244

Saturn

Jupiter

0.004623

Saturn

Tethys

0.00739

Saturn

Dione

0.007904

Saturn

Rhea

0.008511

Saturn

Titan

0.09224

Uranus

Miranda

0.01622

Uranus

Ariel

0.1534

Uranus

Titania

0.07685

Neptune

Triton

1.73

We can see that many of the moons have a significant perturbative influence on the orbital motion of their planets.
It is also interesting to compare the tidal accelerations caused by some of these bodies on others. We will define the tidal acceleration as
the difference between the gravitational acceleration on the near side of the body and that at its center:
accelerationtidal = G * mass1 * (1 / (distance - radius2)2 - 1 / distance2)
where the distance is the distance between the bodies, and the subscripts refer to their labels below. These values have been expressed as
fractions of the tidal acceleration of the Moon on the Earth (1.286 * 10-6 m/s2), and as fractions of the body's own surface gravity.
Here is a time lapse movie of tidal basin filling (17.83 Mb), and one of a tidal bore (6.87 Mb) (the turbulence downstream of a
narrow opening).
In all cases, the length of the orbital semimajor axis has been used to compute the distance:
Body 1

Body 2

atidal (Moon on Earth)

atidal / g

Sun

Mercury

2.955

9.008 * 10-7

Sun

Venus

1.123

1.429 * 10-7

Sun

Earth

0.4481

5.161 * 10-8

Sun

The Moon

0.1218

8.43 * 10-8

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Solar%20System%20Dynamics.html

5/7

1/3/2016

Introduction to Physical Astronomy - Solar System Dynamics

Earth

The Moon

21.7

1.502 * 10-5

The Moon

Earth

1.152 * 10-7

Mars

Phobos

1274

0.387

Mars

Deimos

45.22

0.01899

Jupiter

Io

5476

0.003438

Jupiter

Europa

1167

0.001007

Jupiter

Ganymede

484.5

3.837 * 10-4

Jupiter

Callisto

80.96

7.349 * 10-5

Io

Jupiter

13.33

6.069 * 10-7

Io

Europa

1.081

9.327 * 10-7

Io

Ganymede

0.1029

8.146 * 10-8

Europa

Jupiter

1.593

7.252 * 10-8

Europa

Io

0.6768

4.25 * 10-7

Europa

Ganymede

0.2377

1.882 * 10-7

Ganymede

Jupiter

1.135

5.168 * 10-8

Ganymede

Io

0.1178

7.395 * 10-8

Ganymede

Europa

0.4343

3.748 * 10-7

Callisto

Jupiter

0.1444

6.574 * 10-9

Saturn

Mimas

2067

0.03581

Saturn

Enceladus

1233

0.01818

Saturn

Tethys

1396

0.01058

Saturn

Dione

701.9

0.003384

Saturn

Rhea

351.6

0.001503

Saturn

Titan

95.2

7.935 * 10-5

Saturn

Hyperion

1.837

3.062 * 10-4

Saturn

Iaepetus

1.069

5.825 * 10-6

Tethys

Saturn

0.2478

2.677 * 10-8

Tethys

Enceladus

0.1015

1.496 * 10-6

Dione

Saturn

0.19

2.053 * 10-8

Dione

Tethys

0.1229

9.31 * 10-7

Rhea

Saturn

0.1352

1.461 * 10-8

Titan

Saturn

0.5673

6.13 * 10-8

Titan

Hyperion

0.08113

1.352 * 10-5

Uranus

Miranda

1109

0.01583

Uranus

Ariel

858

0.003604

Uranus

Titania

97.81

2.917 * 10-4

Miranda

Uranus

0.1275

1.626 * 10-8

Ariel

Uranus

0.73

9.309 * 10-8

Ariel

Miranda

0.1663

2.374 * 10-6

Titania

Uranus

0.1409

1.797 * 10-8

Neptune

Triton

368.9

5.33 * 10-4

Triton

Neptune

1.565

1.585 * 10-7

All of the moons in this table are locked in synchronous orbits: their rotational period is equal to their orbital period. This is due to tidal
deformations (bulges toward the planet) which hold the same face toward the planet at all times. Mercury is also in a sychronous orbit, but in
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Solar%20System%20Dynamics.html

6/7

1/3/2016

Introduction to Physical Astronomy - Solar System Dynamics

a more complicated fashion because of its higher eccentricity: it rotates 3 times for every 2 orbits around the Sun.
In addition, it is interesting to compare the tidal accelerations to that of Europa on Io, which is thought to account for much of Io's internal
heating, and subsequent active volcanism. Note that values in the final column are all less than 1. If the tidal acceleration were equal to or
greater than the surface gravity, the body would be gravitationally unstable and would break up. This happens to bodies lying withing the
Roche Limit.
Saturn's Rings lie within Saturn's Roche Limit. (source)

In this image, green hues indicate particle sizes less than 5 cm, while blue indicates less than 1 cm. Purple regions had pieces larger than 5 cm
(up to several meters), and white indicates an area so dense that it was opaque to the radio waves used to probe the particle sizes.
Table of Contents
References
Index
2010, Kenneth R. Koehler. All Rights Reserved. This document may be freely reproduced provided that this copyright notice is included.
Please send comments or suggestions to the author.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Solar%20System%20Dynamics.html

7/7

1/3/2016

Introduction to Physical Astronomy - Some History

Some History
This timeline illustrates the changing rates of progress in our understanding of gravity and planetary motion.

(Click on the timeline to expand it.)


The annotations are roughly separated into "experimental" (above the line) and "theoretical" (below it). Newton is exceptional for appearing on both sides.
The long "dry spell" between Ptolemy and Copernicus is noteworthy: Aristotle's emphasis on the circle as a "perfect" geometric figure seems to have blinded
thinkers for a sizable part of recorded history.
Pythagoras is credited not only with proving the distance theorem, but with realizing that the "evening star" and the "morning star" are one and the
same: Venus.
Aristarchus was the first person to figure out the correct structure of the Solar System.
Ptolemy wrote Almagest, an astronomical treatise based on circular motions of the Sun, planets and stars about the Earth, which codified
astronomical thinking for almost one and a half millenia.
Nicolaus Copernicus (1473-1543) re-invented the heliocentric model of the Solar System, in which the planets revolve around the Sun.
Tycho Brahe (1546-1601) made the most accurate astronomical observations up to his time, without the use of a telescope. Kepler used them to
derive his laws of planetary motion.
Galileo Galilei (1564-1642) performed experiments which led to his development of kinematics: the quantitative description of motion. He built the
first astronomical telescopes, using them to discover the Galilean Moons of Jupiter: Io, Europa, Ganymede and Callisto. He observed Saturn's
rings, although he did not understand their true nature, and he observed the phases of Venus, providing evidence for the validity of Copernicus'
model. His Dialogue Concerning the Two Chief World Systems features a series of illuminating and entertaining discussion between three
characters: Salviati, a proponent of the Copernican view; Simplicio, a proponent of the Ptolemaic view, and Sagredo, who has not yet made up his
mind.
Johannes Kepler (1571-1630) developed his three laws of planetary motion:
1. planetary orbits are ellipses, with the Sun at one focus;
2. orbits sweep out sectors of equal area in equal time, and
3. orbital period2 is proportional to semi-major axis3
Isaac Newton (1643-1727) developed his law of gravitation, designed the reflecting telescope, and used prisms to observe spectra. He also invented
calculus, but we won't be using it much.
Ernest Mach (1838-1916) hypothesized that inertia (what makes it hard to change velocity) is related to the contents of the universe as a whole, an
idea which later influenced Einstein.
Roland Eotvos (1848-1919) performed the first experiments to test the equivalence of gravitational mass (source of the force of gravity) and
inertial mass. This was called the "Equivalence Principle" by Einstein, and is one of the assumptions behind General Relativity.
Albert Einstein (1879-1955) developed the Special (1905) and General (1915) Theories of Relativity. He won his Nobel prize in 1921, however, for
explaining the photoelectric effect, in which photons eject electrons from metals; it took many years for his theories of relativity to gain widespread
understanding and acceptance.
Karl Schwarzschild (1873-1916) discovered the first exact solution to Einstein's equations, later understood to describe a static black hole.
In 1974, Russell Hulse and Joseph Taylor discovered a binary pulsar; after two decades of observations, they determined that its changing rate was
consistent with the gravitational waves predicted by General Relativity. They won the Nobel prize in 1993 for this work.
String Theory, which began in the 1960s as a theory of strong interactions, is an elegant and complex theory which posits that particles are actually
strings. To date it has no experimental verification. Could this be another of Aristotle's circles?
GPS (Global Positioning System) is probably the first example in human experience of the commercialization of so esoteric a theoretical framework
as General Relativity.
(View Cosmos DVD 2, episode 3, Kepler's life.)
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Some%20History.html

1/2

1/3/2016

Introduction to Physical Astronomy - Some History

Some particularly interesting dates:


You need a Java-capable browser to be able to use the applets. If they do not work with your Windows system, download the Java VM (Virtual Machine)
for your version of Windows at the download section at java.sun.com.
These dates represent first successes; for a more complete picture of the history of space exploration, see NASA's Chronology of Lunar and Planetary
Exploration.
Table of Contents
References
Index
2010, Kenneth R. Koehler. All Rights Reserved. This document may be freely reproduced provided that this copyright notice is included.
Please send comments or suggestions to the author.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Some%20History.html

2/2

1/3/2016

Introduction to Physical Astronomy - Spacetime

Spacetime
Our next topic is black holes, but to understand something about them, it is necessary to know about the
concept of spacetime. The idea originated in Einstein's theory of Special Relativity. We will discuss that
part of the theory which deals with the issue of spacetime, but there is much about it that we will leave for
another text...
Special Relativity is a theory of kinematics: it gives us a framework for describing the world, but it does not
predict how the world will behave under any given conditions. What makes SR special is that it forces us to
treat time as a relative of space: there is no separate space which we measure with meter sticks, and a
separate time which we measure with clocks. There is only spacetime, and every measurement requires both
a meter stick and a clock. And this presents a problem.
We are familiar with three spatial directions: you may think of them as "left-right", "up-down" and "in-out". A
mathematician would perhaps give those directions labels; for instance, "x", "y" and "z". Our whole biology,
and most especially our minds, are "wired" to understand life in "3-D": three spatial dimensions. But if time is
an integral part of spacetime, and cannot be consistently divorced from it, spacetime must be "4-D": it must
have four dimensions. And nothing in our biology is fit to understand four dimensions.
Consider a planet in a circular orbit around a star. We observe its motion one moment at a time; we can only
"see" its previous motion in our memory, and we can only anticipate its future motion if we have studied hard
enough. But our sensory experience occurs as a sequence of spatial observations:

A being who could see time in the same context as the spatial dimensions would see something entirely
different. Since we know our orbiting planet is essentially motion in a plane, we can show you what this
superior being observes by plotting time along the axis perpendicular to the plane of the orbit:

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Spacetime.html

1/6

1/3/2016

Introduction to Physical Astronomy - Spacetime

The "cylinder" along the time axis is the continuous view of the star, as the planet "spirals" in the "plane-time"
view.
Of course, when the motion is truly three-dimensional, we cannot construct a static image of its movement in
space and time. A mathematician would say that the embedding of a four-dimensional object in three
dimensions is impossible. But all is not hopeless: there are tools which can help us. One such tool is the
projection.

Projections
If you have ever taken a picture, you are already familiar with the idea of a projection. These photos of the
famous Japanese movie idol Gojira are projections : a mapping of three dimensions into the flat twodimensional images you see.
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows
system, download the Java VM (Virtual Machine) for your version of Windows at the download section at
java.sun.com.
Whenever you take a picture, you are projecting a 3-D object into 2-D, and in the process you are ignoring
the information in the third dimension. We are of course assuming here that you are not taking stereoscopic
pictures, which allow your eyes to reconstruct the original 3-D object from a pair of 2-D images taken from
slightly different positions.
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Spacetime.html

2/6

1/3/2016

Introduction to Physical Astronomy - Spacetime

But these photos have actually projected out another dimension: time. In them we see Gojira posing for the
cameras of the paparazzi, but we have lost all of the information about what cities he may have demolished
before or after the photos were taken. Every photograph is a 2-D projection of a 4-D object: it throws
away information about an object that occupies three spatial dimensions and one temporal (timelike)
dimension by depicting only two of the spatial directions, and freezing the object in an instant of time.

Spacetime
Spacetime is made up of events. Each event is specified by its spatial coordinates and the time at which it
occurred. The collection of all the events in the life of an object is called its world line. Since physicists
cannot embed 4-D spacetime in 3-D or 2-D any more than you can, they often choose to project spacetime
into two dimensions, one spatial and one temporal, and plot the world line of an object as a simple path on a
plane.
Since we often are concerned with the distances between objects, we often take the horizontal axis of our
plot to be the distance from the origin of our coordinate system, and the vertical axis to be the time. In
addition, we often take t=0 to be some significant event. The fact that some times are therefore negative
simply indicates that the events at those times occurred before that significant event.
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows
system, download the Java VM (Virtual Machine) for your version of Windows at the download section at
java.sun.com.
Since the radius of a circle is a constant, the world line of the planet in the r-t plane is a straight vertical line.
Both plots ignore any angular information about the respective world lines: we do not know in which
direction either object lies at any point in time. That information has been lost in the projection. But in many
cases the system has sufficient symmetry that the angular information is irrelevant. If we take either of these
systems as isolated (perhaps hidden in a vast dust cloud which renders the rest of the universe invisible), the
only preferred direction is the direction defined by the axis of rotation of the star. If we take that direction
to be along the z axis (as is customary), the only piece of information missing in either plot is the inclination of
the orbit: how tilted it is with respect to the z axis. If the cometary system is not isolated, we are also missing
information about the direction of its orbital axis with respect to, say, the direction toward the center of the
galaxy. But in many cases of interest, there is sufficient symmetry for a 2-D plot of world lines to be
essentially complete.

Light Cones
There is one small detail we have overlooked in the second plot: the scale between the r and t axes. Distance
is measured in, for instance, meters. Time is typically measured in seconds. In order to plot world lines
consistently, we must set the scale between space and time. But the universe has done that for us, in providing
something which moves at the same speed no matter what coordinate system you are using: light. Light
always moves at 299,792,458 meters each second (almost 675 million miles an hour), no matter which
frame of reference (coordinate system) you adopt.
Using coordinates (r,ct), light moves along lines which form 45 degree diagonals with the axes.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Spacetime.html

3/6

1/3/2016

Introduction to Physical Astronomy - Spacetime

In this plot, the line going from lower left to upper right represents a ray of light moving away from the origin,
while the line going from lower right to upper left represents a ray of light moving toward the origin. We call
the region between those lines the light cone of the event at t=0 (where the light rays crossed). The (shaded)
lower part is called the past light cone of the event and the (shaded) upper part is called the future light
cone of the event.
The past light cone includes all events that could have influenced the event at t=0, and the future light cone
includes all events that the event at t=0 could influence. Note that every event (every point on the r-t plane)
has its own light cone.
The light cone of a planet at one event on its world line.

If we plot the comet's world line with our new scale, it looks identical to the planet's. Its orbital motions are
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Spacetime.html

4/6

1/3/2016

Introduction to Physical Astronomy - Spacetime

obscured because the speed of light is so fast that the scale has flattened out the variations in its orbital
distances.
Two world lines, with their respective light cones at one instant of time.

Now the plot thickens: at the instant of time in which both sets of cones were drawn, there was no way for
these two objects to communicate. Each was outside the other's light cone, and so no signals could have
been sent from one to the other. This implies that there was no way that they could have known that this
snapshot was taken at one instant of time: there is in general no such thing as simultaneity between any two
observers. In addition, this plot illustrates the notion of horizons: the observer on the right can only be
affected by the observer on the left in regions where their light cones overlap, so at the time these light cones
were drawn, they were outside each other's particle horizon. We also say that they were separated by a
spacelike interval.
In any plot of world lines, if two events are connected by a line shallower than 45 degrees, the interval
between them is spacelike. If they are connected by a line steeper than 45 degrees, they are separated by a
timelike interval: one's past light cone is within the other's future light cone, and so communication between
them along that interval was possible. In the graph above, the upper green region is the region of either world
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Spacetime.html

5/6

1/3/2016

Introduction to Physical Astronomy - Spacetime

line which can be affected by the event of the other which occurred at t=0 (and whose light cones were
drawn in the graph). The lower green region is the region of either world line that could have had an affect on
the event of the other which unfolded at t=0.
So if the observer on the right sent a radio message to the observer on the left at the point marked "a", the
left-hand observer would receives that signal at point "b", and the return message would be received by the
right-hand observer at point "c".

The Fine Print


The speed of light has been measured carefully (and continues to be measured more and more carefully as
experimental techniques are improved), and it is found to have exactly the same value in every reference
frame. We are not equipped to prove so here, but that indicates that Special Relativity is a fact: whether we
like it or not, we live in a 4-D spacetime. But so long as we work with velocities that are small compared to
the speed of light, we can ignore this fact and treat space and time as fundamentally distinct concepts.
When velocities are close to that of light, formulas like the Pythagorean Theorem no longer hold and must be
modified to take Special Relativity into account. This causes observers moving at relativistic speeds with
respect to other observers to measure length and time differently: they will not agree on how long a meter is,
or what happens in a minute.
Finally, we feel compelled to remind you that in each of the 2-D world line plots, each point in the r-t plane
represents the surface of a sphere, since all directional information has been projected out. Try not to think
too hard on that until the rest of this seems natural to you. If you do think about it before you are ready, you
will get a headache. Trust us.
Table of Contents
References
Index
2013, Kenneth R. Koehler. All Rights Reserved. This document may be freely reproduced provided that
this copyright notice is included.
Please send comments or suggestions to the author.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Spacetime.html

6/6

1/3/2016

Introduction to Physical Astronomy - Spectra

Spectra
We are all familiar with the spectrum of colors hidden in sunlight, revealed by a prism. But we need to be more specific, and
distinguish between three tyes of spectra:
continuous spectra,
emission spectra and
absorption spectra.
Continuous spectra is that rainbow of colors displayed by sunlight passing through a prism. It results from the black body thermal
radiation generated deep within the Sun:
Black body curves for Sirius A (9200 K), Sol (5800 K) and Betelgeuse (3800 K) (dashed curves are proportional to actual
intensities).

The location of the peak and overall shape of the curve are determined by the the surface temperature of the star in Kelvin (=
Celsius + 273.15).
Emission spectra are the bright lines generated by electron transitions. Absorption spectra are dark lines in the continuous
spectrum caused when an atom or molecule absorbs a photon, and one of its electrons jumps to a higher energy level. Absorption
lines are widely used in astrophysics, allowing the investigation of chemistry on other worlds and chemical makeup and nuclear
processes in stars from our own Sun to those literally across the universe.

The Spectrum Viewer


The following applet can supply digital spectra, along with images, parallax (when known) and radiation flux information, for 46 stars
covering the seven major stellar types. It also knows 1226 spectral lines associated with 34 elements, ions and molecules commonly
found in stellar atmospheres. It allows you to fit a black body distribution to the spectrum, to investigate portions of the spectra in
detail, and it allows you to fit a Gaussian distribution ("bell curve") to individual lines to quantify line broadening.
The spectral data varies in resolution and coverage among the stars: some spectra include ultraviolet and/or infrared wavelengths
while others cover only part of the optical spectrum. All are flux-calibrated spectra, meaning that the flux values have calibrated
physical units of Watts per square meter per Angstrom. Some of the original spectra were compiled at resolutions greater than the
applet's finest resolution of one value per Angstrom, and those spectra have been smoothed for our use here. In addition, some
spectra have been truncated because of the applet's limited color resolution: it can show only 256 levels of intensity for any given
hue.
Begin by choosing a star from the pop-up menu. An image of the star will be displayed and, for most stars, the measured annual
parallax angle and absolute error in that measurement will be displayed in the text box to the right of the image.
Note that the image is not, except in the case of Sol, an image of the stellar surface. It is actually a diffraction pattern
(interference pattern) with a strong, wide central maximum (the central circular "image") and in most cases indiscernible
secondary maxima. In most cases, diffraction spikes will be visible; these indicate that the star is relatively close to
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Spectra.html

1/8

1/3/2016

Introduction to Physical Astronomy - Spectra

us.
For all stars, the text box will also contain the minimum and maximum flux values over the range of the displayed spectrum, along
with the total flux over that range.
Below this is another pop-up menu which selects an element (or ion or molecule) for spectral line identification, followed by a button
and scroll bar enabling a black body temperature fit to the spectrum. Beneath those are three displays. The top display shows the
spectral lines associated with the element selected (they are displayed as emission lines, that is, bright against a black background).
Below that is a reconstructed color image of the actual spectra (wavelengths in the ultraviolet and infrared are displayed in shades of
gray). Finally there is a graph of the flux as a function of wavelength:
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows system, download the
Java VM (Virtual Machine) for your version of Windows at the download section at java.sun.com.
You may click on either the spectrum image or graph to mark a central wavelength value on which to focus for detailed inquiries.
The "Zoom In" and "Zoom Out" buttons allow you to quickly change the range of wavelengths displayed about that central value,
and the values of the wavelength at the left and right edges of the graphs are displayed to the right of those buttons. The scroll bar
labeled "Central lambda" allows you to smoothly vary the central value, and the value of the flux at the central wavelength is
displayed to the right of the central lambda value. Finally, the "Gaussian Fit" button draws a Gaussian fit to the spectrum graph
centered on the central lambda value, with a standard deviation (half-width) as specified by the scroll bar to the right.
(source, source, source, source, source, source, source)

Absorption Lines and Temperature


The Spectrum Viewer will allow you to analyze spectra of several stars in detail. Using the energy flux and parallax data it provides,
we can compute a star's luminosity using the equation
energy flux = luminosity / (4 distance2)
Using values from the Spectrum Viewer for the star 108 Virginis, we see from its parallax (5.31 mas) that its distance from Earth is
r = 1000 pc/mas / 5.31 mas = 188.3 pc, times 3.086 * 1016 m/pc
= 5.81 * 1018 m
The Viewer also tells us that its total flux is 9.9775 * 10-11 W/m2, so its luminosity is
L = 4 * (5.81 * 1018 m)2 * 9.9775 * 10-11 W/m2
= 4.23 * 1028 W, divided by 3.84 * 1026 W for our Sun
= 110 Lsolar
Note that we do not have a "complete" spectrum; there appears to be additional flux in the ultraviolet that is missing from the data
for this star. Because of this, we have really underestimated the luminosity of this star.
We can find the effective temperature of a star as follows. Use the "Black Body" button and adjust the temperature scroll bar so that
the black body curve is as close to the spectrum as possible. Since a star is not a perfect black body, we have some choices here:
we can match the peak wavelength (in this example, around 3970 Angstroms, giving us a temperature of 7300 K) or we can match
the leading or trailing slope (in this example, closer to 9300 K). In general, matching the trailing slope tends to be more accurate
than the other options.
A given atom, ion or molecule only exists in a star's atmosphere for a range of temperatures. At lower temperatures, an ion may
capture one or more free electrons to become neutral. At higher temperatures, collisions may cause an atom to lose one or more
electrons to become ionized. In either case, this will change the wavelengths of its absorption lines. At sufficiently high temperatures,
molecules will break up into their constituent atoms, and their lines (or bands) disappear altogether. We can use the following table
to look for some characteristic lines as a check on our black body temperature. Choose the element from the drop-down menu
("Reference Spectra Element") and look for consistent matches for the lines displayed, indicating the presence of that element in the
star's atmosphere. If the lines in the table are present, the temperature should be in or close to the range given, and if the lines are
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Spectra.html

2/8

1/3/2016

Introduction to Physical Astronomy - Spectra

extremely strong (dark), the temperature should be close to the peak temperature. The wavelengths for the CH and TiO bands are
the wavelengths at the left-hand edge of the bands.
specie

lines (Angstroms)

T Range (K)

Peak T (K)

3966, 4097, 4336, 4856, 6555

5000-40000

9000

He

4471, 4542

10000-50000

29000

Ca

4227

2000-5500

3000

Ca+

3934, 3968

3000-7000

5000

Na

5890, 5896

2000-5500

3000

Mg+

2796, 2802, 4481, 5173

8000-30000

9200

Si+

4128, 4131

8000-20000

9200

Si++

4552

20000-40000

25000

Si+++

1394, 1403, 4089, 4116

30000-50000

40000

Fe

4045, 4143, 4299, 4325, 4384, 5270

2000-7000

4500

Fe+

4173, 5316

4000-8000

5700

CH band

4300-4315

5000-6000

5500

TiO bands

4762, 4955, 5167, 5448, 5862, 6159, 6384

2000-4000

3000

Here are sample optical spectra arranged by spectral class; can you identify the prominent lines?
Sample optical stellar spectra. (source)

The spectrum for 108 Virginis is a low resolution spectrum, so all we can see are the broad Hydrogen lines, which do not help much
in verifying our black body temperature. This check is much more useful with high resolution spectra, especially spectra in the visible
and near ultraviolet.
With the temperature and the luminosity, we can compute the star's radius using
luminosity = 4 radius2 temperature4
Using our temperature of 9300 K for 108 Virginis (and 5.67 * 10-8 for ), we obtain
r = (4.23 * 1028 / 4 93004)1/2

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Spectra.html

3/8

1/3/2016

Introduction to Physical Astronomy - Spectra

= 2.82 * 109 m, divided by the radius of our Sun (7 * 108 m)


= 4 rsolar
Since our estimate for the luminosity was low, this is an underestimate for the radius. But because of the square root, we are closer
here than in our luminosity calculation (if the luminosity was off by a factor of 2, the radius will only be off by a factor of 21/2 =
1.414).
Note that the presence of metallic absorption lines indicates that the star is a later-generation star: the metals were inherited from
earlier-generation stars.
Portfolio Exercise: You will be given a list of 9 stars which you are to analyze for several portfolio exercises. In this
exercise,
compute each star's distance in parsecs and in meters;
compute each star's luminosity in both Watts and solar units;
find the black body temperature for each star; and
compute each star's radius, both in meters and in solar units.
For one or more of your stars, the applet may not have parallax information; for these you will only be able to find the
black body temperature. Several of the spectra show strong emission lines (associated with flares) which alter the scale
of the graph and make black body temperature estimation more difficult.

Doppler Shift
Line of sight motion can be computed using the shift in the central wavelength of well-known absorption lines:
apparent wavelength / actual wavelength = 1 + recession velocity / c
If the apparent wavelength is shorter than the actual, the motion is toward us and the recession velocity is taken to be negative.
A dramatic illustration of the Doppler effect provided some of the first real evidence for the existence of black holes. Since
recessional motion (away from us) causes the apparent wavelength to be larger, color is shifted toward the red end of the spectrum;
this is called red shift. If the source is moving toward us, the apparent wavelength will be smaller, and color will be shifted toward
the blue end of the spectrum: blue shift. When red and blue shifted light comes from adjacent areas in a very small region of the
center of a galaxy, it is a signal that matter is rotating very quickly around a compact object. In this case, the only type of object
expected to be able to cause rotational motion of that magnitude is a black hole.
Red and blue shift in the nucleus of M 84 signals the presence of a black hole. (source)

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Spectra.html

4/8

1/3/2016

Introduction to Physical Astronomy - Spectra

The Spectrum Viewer does not have sufficient resolution to enable us to compute recession shifts for any of the stars it has data for.
But we will return to this computation when we discuss Cosmology.
The broadening of most narrow absorption lines is primarily due to the motion of the atoms involved, although pressure and other
factors can be equally important. For instance, Hydrogen and Helium lines are broadened by the Stark Effect: since their nuclear
charge is small, they are particularly susceptible to electric fields created by ions in the stellar atmosphere. These fields perturb their
energy levels and are primarily responsible for their line broadening.
Atomic speeds in the photosphere can be estimated using the width of some absorption lines:
line width = 2 * speed * central wavelength / c
The line width is twice the standard deviation reported by the Spectrum Viewer.
The spectrum for 108 Virginis has too low a resolution to be able to compute Doppler Shifts; only the broad Hydrogen lines are
present, and these are unsatisfactory for computing atomic velocities, because of the Stark Effect. But using the prominent Sodium
line of Epsilon Indi (at 5890 Angstroms), we can compute an approximate speed for these atoms in the star's atmosphere as
follows. Click on the graph at 5890 Angstroms, and "Zoom In"; if the central wavelength is not 5890, click again or use the "Central
Lambda" scroll bar until it is. Then click the "Gaussian Fit" button; a small curve appears around the line. Change the Standard
Deviation using the scroll bar until the curve is a close fit to the line on the graph. In this example, it is a close fit between 1.5 and
2.4. Taking the average (1.95), we compute the approximate speed of these atoms as
v = 1.95 * 3 * 108 / 5890 = 99000 m/s
Portfolio Exercise: Using both the Calcium lines near 3900 Angstroms, and the Sodium lines near 5900 Angstroms,
compute the approximate atomic speeds from the line width, for each star.
If the lines are not clearly defined and relatively deep, it will not be possible to do this. This will occur if
one or more of your stars have low-resolution spectra, or those lines are located in very low-flux
wavelengths. List those stars which you are unable to evaluate.
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Spectra.html

5/8

1/3/2016

Introduction to Physical Astronomy - Spectra

Are the speeds consistent for each star? What does the range tell you about this computation?

Extinction
Taking inventory of our equations involving luminosity and magnitude, we see that we have three equations in five unknowns:
energy flux = luminosity / (4 distance2)
M = m + 5 - 5 * log10 distancepc
m = 4.83 - 2.5 * log10 luminositysolar
energy flux, luminosity, distance, absolute magnitude (M) and apparent magnitude (m). So in principle, if we know any two of these
quantities, we can use the three equations to compute the others. Photometry gives us the energy flux and apparent magnitude, so
we seem to be home free. But there is an assumption that we have ignored, and which we must take into account: space is not
entirely empty.
The volume in the immediate neighborhood of the Earth has a mean density of about 5 atoms per cubic centimeter (mostly
Hydrogen (source); 1 cm3 = 10-6 m3). This falls to about 0.3 in the Solar neighborhood and to less than 0.001 in the "Local
Bubble":
Our Sun's neighborhood in the Milky Way. (source)

In interstellar space the average density is around 1 atom per cm3, and it is even less in intergalactic space. Larger dust particles are
about a million times rarer, except in nebulae and clouds, where the density runs between 10 to a million atoms, molecules or dust
particles per cm3. But since space is so vast, some of the electromagnetic radiation traveling through the interstellar medium is
absorbed, and because shorter wavelengths are scattered more than longer ones, there is additional attenuation in the red and
infrared wavelengths. The overall loss is called extinction, and the preferential scattering causes reddening. The following
absorption nebula illustrates both phenomena:
Molecular Cloud Barnard 68, showing both extinction and reddening. (source)

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Spectra.html

6/8

1/3/2016

Introduction to Physical Astronomy - Spectra

While the spectra in the Spectrum Viewer have been corrected for extinction and reddening, it is important to know how
astronomers can compute the necessary corrections. Armed with an independent measure of distance (for instance, from the
parallax), we can compare the observed apparent magnitude with the expected apparent magnitude. The result is
extinction = (mobserved - mexpected) / distance
here measured in magnitudes per kiloparsec; note that it is a function of wavelength.
Spectral classification is particularly useful here. Along the main sequence of a Hertzsprung-Russell diagram, there is essentially a 1:1
correspondence between spectral class and absolute magnitude. Because of this, stars of a given spectral class but at various
distances can be used to compute extinction. For instance, the star BD +31 640 is located in the open galactic cluster IC 348. It is
an A3V star located 236 parsecs away, with an absolute magnitude of 1.5. This means the expected apparent magnitude (m = M 5 + 5 log D) is
1.5 - 5 + 5 log 236 = 8.36
However, the observed apparent magnitude (measured near 5448 Angstroms, the effective wavelength for the V band) is 11.4
(source), meaning that the extinction is
(11.4 - 8.36) / 0.236 = 12.86 magnitudes / kpc
In this case, the extinction is due largely to an obvious cloud:

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Spectra.html

7/8

1/3/2016

Introduction to Physical Astronomy - Spectra

and it is more reasonable to quote the extinction simply as 3.04 magnitudes.


Table of Contents
References
Index
2010, Kenneth R. Koehler. All Rights Reserved. This document may be freely reproduced provided that this copyright notice is
included.
Please send comments or suggestions to the author.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Spectra.html

8/8

1/3/2016

Introduction to Physical Astronomy - Stellar Evolution

Stellar Evolution
Stellar Evolution (I)
In general,
stellar lifetimeyears = 1010 masssolar / luminositysolar
Since luminosity is approximately equal to mass3.5 (in solar units), this is equivalent to
stellar lifetimeyears = 1010 / masssolar2.5
leading us to the conclusion that more massive stars burn out more quickly.
We will continue to use the common term "burning" for the nuclear reactions which power stars. But it is
important to understand that stars are not "on fire": there is no organic material to be oxidized, and not much
Oxygen to do the oxidation.
These graphs illustrate the lifetimes of stars of 0.5, 1, 1.5, 3, 5 and 9 solar masses, from various perspectives (use the Shift
key to expand them):
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows system, download
the Java VM (Virtual Machine) for your version of Windows at the download section at java.sun.com.
The brown phase represents the Hyashi Track, during which the protostar shrinks significantly, dropping in
luminosity but increasing in temperature due to the increased internal pressure. At the beginning of the Hyashi Track,
the core is convective. For stars less than 1.3 solar masses, the core is radiative by the end of the track, and the pp
chain dominates the fusion reactions that begin then. For stars greater than 3 solar masses, the core is convective
throughout and the CNO cycle dominates.
The green phase represents the approach to equilibrium that precedes the main sequence. During this phase, the
outward pressures from the fusion reactions in the core have not come into equilibrium with the gravitational
pressures tending to contract the star, resulting in winds significantly larger than those generated by a main sequence
star.
The time of the beginning of the main sequence (shown in yellow) is called the Zero Age Main Sequence. A star
arrives on the main sequence then, and stays essentially at that one place on the H-R diagram as long as it continues
to fuse Hydrogen into Helium in its core.
What happens next is a tug of war between the opposing pressures due to fusion and gravity: as fusion
products build up in the core, its density increases. The core shrinks, the pressure increases, the
temperature increases, heavier nuclei begin to fuse in the core and previous fusion reactions move
outward into shells. As the fusion reactions take place in regions closer to the surface of the star, the
outward pressure from those reactions enlarge the star. Depending on the size of the star, this process
can repeat, with multiple shells supporting different reactions simultaneously.
The period of Hydrogen shell burning is shown in orange, and the red giant branch is shown in red. During this
portion of stellar evolution, the star's luminosity and radius increase enormously, but the expansion is cooling and
results in a drop in surface temperature.
At the end of the red giant branch, if the star is less than 0.5 solar masses, no Helium fusion occurs. If the star is less
than 2.25 solar masses, the Helium core has already begun to collapse into a degenerate state, and there is a Helium
flash which begins Helium fusion. Larger stars begin burning Helium in a smooth transition.
The change in stellar radii as from the beginning of the main sequence to the end of the red giant branch; the black
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Stellar%20Evolution.html

1/8

1/3/2016

Introduction to Physical Astronomy - Stellar Evolution

vertical line indicates 1 A.U; the vertical axis is mass.

The horizontal branch is shown in purple. This is another equilibrium state, during which the star's luminosity is
almost constant. Its surface temperature and radius may change depending on the nuclear reactions underway. For
stars greater than 4 solar masses, Carbon fusion also occurs during the horizontal branch (and a Carbon flash can
occur). For stars greater than 8 solar masses, Oxygen fusion also occurs, and for stars larger than 10 solar masses,
Silicon burning occurs in the terminal stages before supernova.
The fusion of successively heavier nuclei is an accelerating process, due to the drop in the number of nuclei available
to fuse. Carbon fusion might last for hundreds of years before Oxygen fusion begins, but Oxygen burning might last
only a year or so before Silicon fusion begins. It takes about a day for the core to fuse into Iron.
By examining an H-R diagram for the stars in a globular cluster like M 55, it is possible to determine the age of the cluster:
H-R diagram for M 55. (source)

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Stellar%20Evolution.html

2/8

1/3/2016

Introduction to Physical Astronomy - Stellar Evolution

Since more massive stars have shorter lives, the "end" of the main sequence tells us how old the cluster is. Compare visible
and ultraviolet images of the cluster Omega Centauri; note the obvious population differences between hot and cooler stars:
Omega Centauri in visible and UV. (source)

Numerical values in this section are approximate. Stellar life cycles are based on models described in 1965
and 1967 by Iben (Astrophysical Journal 141:993 and Annual Reviews of Astronomy and Astrophysics
5:571, respectively).

Stellar Evolution (II)


White Dwarfs
The final fate of a star depends on its mass at the termination of fusion processes. Stars born with less than 8 solar masses
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Stellar%20Evolution.html

3/8

1/3/2016

Introduction to Physical Astronomy - Stellar Evolution

experience a series of Helium shell flashes which cause the luminosity to fluctuate and the outer layers to be ejected, forming
a planetary nebula. These stars typically shed enough mass so that their final mass is less than the Chandrasekhar Limit
of 1.44 solar masses to become a white dwarf: the compressed core of the final fusion product.
Note the shapes of the nebulae in these images; the processes which create them must be far less symmetrical than the naive
spherical stellar model which all our discussion of radii imply:
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows system, download
the Java VM (Virtual Machine) for your version of Windows at the download section at java.sun.com.
A white dwarf is a degenerate star because its atoms are as close as possible within the constraints of the Pauli Exclusion
Principle, which tells us that no two electrons can be in the same state. In this condition, every atom is in the lowest state
possible, so that the number of available states, or degeneracy, is as low as possible.
Sirius B is a white dwarf whose absolute magnitude is 11.2 and whose mass is approximately equal to that of our Sun. Its
luminosity is then
L = 10(4.83 - M)/2.5
= 10-2.548 = 0.0028 Lsolar
Its black body temperature is around 14,800 K, so its radius is
r = (L / (4 T4))1/2
= (0.0028 * 3.84 * 1026 / (4 148004))1/2
= 5.61 * 106 m = 0.008 Rsolar
From this we can compute its average density
= m / (4/3 r3)
= 2 * 1030 / (4/3 (5.61 * 106)3), times 0.001 to convert kg/m3 to g/cm3,
= 2.7 * 106 g / cm3 (!)
and its escape velocity
escape velocity = (2 * G * mass / radius)1/2
= (2 * G * 2 * 1030 / 5.61 * 106)1/2
= 6.88 * 106 m/s
= 0.023 c.
Its surface gravity is
surface gravity = G * mass / radius2
= G * 2 * 1030 / (5.61 * 106)2
= 4.22 * 106 m / s2,
which means that if you could survive on its surface, you would weight about 430 thousand times as much as you do on
Earth! These values are typical orders of magnitude for white dwarfs:
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Stellar%20Evolution.html

4/8

1/3/2016

Introduction to Physical Astronomy - Stellar Evolution

White Dwarf

Mass solar

Absolute Magnitude

Radius solar

40 Eridani B

11

0.501

0.0136

Procyon B

13.2

0.604

0.0096

Sirius B

11.2

0.0084

(source)
Note that at 8.6 light years, Sirius B is the closest white dwarf to Earth. That, coupled with the fact that it is a binary partner
to Sirius A, allows us to understand it much better than many other white dwarfs. Careful observation of their orbital period
and separation was the key to computing its mass.

Neutron Stars
Stars born larger than 8 solar masses usually retain enough mass to undergo core collapse, with the resulting shock wave
producing a Type Ib supernova (spectra without Hydrogen or Silicon lines, with Helium lines), a Type Ic supernova
(without Hydrogen or Helium or Silicon lines) or a Type II supernova (with Hydrogen lines). The presence of specific
lines in their spectra probably indicates which outer layers were blown off before the explosion occurred. These supernova
types can also be classified by their light curves, which show their energy output (here in the blue band) as a function of
time:
Supernova light curves.

For Type II supernovae, the "-P" and "-L" designations stand for "plateau" and "linear". The timings of the light curve decays
seem correlated with the radioactive decay rates of 56Ni and 56Ni.
If the final mass at core collapse (after the shells have been shed) is less than 3-3.2* solar masses, the remnant will be a
neutron star. In a neutron star, the atoms have come so close to each other that the electrons have been absorbed into the
nuclei, releasing neutrinos. The remaining neutrons are now degenerate: they are as close as possible within the constraints
of the Pauli Exclusion Principle. A neutron star is essentially one large, extremely dense nucleus, with no net electric charge.
*Most neutron star masses are close to 1.4 Msolar; the largest measured so far is about 2.1 Msolar. The actual
limit depends on how neutron matter behaves, and is the subject of ongoing research.
If the neutron star is rotating, and has a sufficiently large magnetic field whose axis is not along the rotation axis, we may
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Stellar%20Evolution.html

5/8

1/3/2016

Introduction to Physical Astronomy - Stellar Evolution

observe it as a pulsar. It is assumed that most if not all stars rotate, so by conservation of angular momentum, a collapsed
core such as a white dwarf or neutron star will rotate much faster than its progenitor, by a factor of
(r before / r after)2.
The neutron star at the heart of the Crab Nebula has about 1.4 times the mass of Sirius B but its radius is only about 10
kilometers. This is 560 times smaller, so its density is 250 million times larger, its escape velocity is 28 times larger (0.64 c)
and its surface gravity is 440 thousand times greater! We have already seen the Crab Nebula in multiple wavelengths; here
is an x-ray image of the neutron star and supernova remnant Cassiopeia A:
Neutron star and supernova remnant Cassiopeia A, in x-rays. (source)

Planetary nebulas glow from irradiation by the remnant white dwarf or neutron star.
Now consider our old friend Betelgeuse. As we mentioned in previously, it is an M2Ib star located 427.3 light years from
us. Its temperature is about 4000 K, its mass is about 20 Msolar and its luminosity is about 50,000 Lsolar. That means that
its radius is
r = (50000 * 3.84 *1026 / (4 40004))1/2
= 3.25 * 1011 m = 464 Rsolar
Since its luminosity is varying, we know it is in the horizontal branch, and it is estimated that within 10,000 years it will
explode as a supernova. Assuming the supernova has a peak luminosity of 1010 Lsolar, the energy flux we should
experience here on Earth is
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Stellar%20Evolution.html

6/8

1/3/2016

Introduction to Physical Astronomy - Stellar Evolution

energy flux = luminosity / (4 distance2)


= 1010 * 3.84 *1026 W / (4 (427.3 ly * 9.461 * 1015 m/ly)2)
= 0.0187 W/m2
This may not seem like much compared to the total flux we receive from the Sun, but consider that much of this is in gamma
radiation. If we assume the normal gamma flux is 5 MeV/cm2/s, or around 8 * 10-9 W/m2, this is a lot of radiation. The
lesson here is that the neighborhood affected by a supernova is very large. In any given galaxy, every 30 to 50 years
another star explodes as a supernova:
SN 1994D in NGC 4526. (source)

This makes the universe seem like a violent place, and it can be. But it is important to remember that the explosion of a
supernova not only disrupts its neighborhood, but it also seeds the interstellar medium with heavier elements, and the shock
waves compress that medium so that more stars can develop in the future:
Henize 206. (source)

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Stellar%20Evolution.html

7/8

1/3/2016

Introduction to Physical Astronomy - Stellar Evolution

Star forming regions in galaxies are identified by the hydrogen gas glowing in infrared and by the ultraviolet light of young,
bright stars.
Table of Contents
References
Index
2010, Kenneth R. Koehler. All Rights Reserved. This document may be freely reproduced provided that this copyright
notice is included.
Please send comments or suggestions to the author.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Stellar%20Evolution.html

8/8

1/3/2016

Introduction to Physical Astronomy - Stellar Populations

Stellar Populations
Here we are interested in all stars, and in classifying them as an aid to understanding the physics behind them.

Spectral Classes and Effective Temperatures


Spectral classes were originally defined using visual cues from spectra. We will define them in terms of effective surface
temperature, using the letters "O", "B", "A", "F", "G", "K" and "M" (from hottest to coolest) and subdivide each main class into 10
subclasses, labeled by the digits 0 through 9 (again, hottest to coolest). The following plot gives the most likely effective temperature
for each spectral class:
Effective temperature vs. spectral class; red/orange = main sequence, red = giants, blue = supergiants.

Using our temperature of 9300 K from the Spectrum Viewer, we see that 108 Virginis is likely close to a class A1 (interpolating
between the temperatures for A0 and A2 in the graph).

Magnitude, Luminosity Classes and Spectral Classes


If we know the luminosity, or the apparent magnitude and the distance to the star, we can compute the absolute magnitude using one
of the equations
absolute magnitude = 4.83 - 2.5 * log10 luminositysolar
absolute magnitude = apparent magnitude + 5 - 5 * log10 distancepc
For 108 Virginis, at a distance of 188.3 parsecs and a luminosity of 110 Lsolar, we obtain an absolute magnitude
M = 4.83 - 2.5 * log10 110 = - 0.27
and an apparent magnitude of
m = - 0.27 - 5 + 5 * log 10 188.3 = 6.1
Note that because of our underestimate of the luminosity, M is larger (less negative) than it should be, and m is also larger than it
should be. Here the error is even less than in our radius calculation because the log function increases more slowly than the square
root.
Standard luminosity classes are
Ia - Bright Supergiants
Ib - Supergiants
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Stellar%20Populations.html

1/5

1/3/2016

Introduction to Physical Astronomy - Stellar Populations

II - Bright Giants
III - Giants
IV - Subgiants
V - Dwarfs (including Main Sequence Stars)
A spectral class beginning with "D" indicates a degenerate star, or white dwarf: a star which has collapsed to a small,
hot cinder after fusion processes have stopped. Additional designations include "e" (has emission lines), "n" (has diffuse
lines), "p" (has a peculiar spectrum; even physicists don't understand it all yet!), "s" (has sharp lines), and "v" (variable
star).
The following table relates the absolute magnitude to the luminosity class for each spectral class (the magnitudes given are the most
likely values for each class):
Absolute magnitude vs. spectral class; orange = main sequence, red = giants, blue = supergiants, green = white dwarfs.

Using the estimated class A1 for 108 Virginis, we see that it could conceivably be a main sequence star, a supergiant, or a white
dwarf. Given an absolute magnitude of -0.27, it is much more likely that it is a main sequence star.
Using SIMBAD, you can look up the accepted spectral class for 108 Virginis. You will find it listed as "B9.5V". Our choice of A1
is very close to B9.5, and the "V" indicates a main sequence star, so our choice of luminosity class was correct.
Portfolio Exercise: Using the information you have collected so far, assign a spectral class to each of your stars from
the spectra portfolio exercises.
For each star, compute the absolute and apparent magnitudes. Assign a luminosity class to each star.
Look each star up on SIMBAD. Comment on your success rate.

The Hertzsprung-Russell Diagram


Suppose you could take a picture of every person on Earth at a given instant of time. Armed only with that picture, you must find a
model that explains the human life cycle, from conception to death. The Hertzsprung-Russell Diagram is like that picture, and
with it we can learn some things about the lives of stars. Note that it is like a mirror image of the graph above:
A Hertzsprung-Russell Diagram. (source)

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Stellar%20Populations.html

2/5

1/3/2016

Introduction to Physical Astronomy - Stellar Populations

The position of a star on the main sequence (the line from upper left to lower right) is determined by its mass. Main
sequence stars of spectral type G0 are about 1 Msolar, while those of type B0 are about 15 Msolar (stars begin their
main sequence lives with masses varying from 0.07 to about 200 Msolar (source)). Almost half of the stars in the
neighborhood of our Sun have a mass less than about a third of the Sun's; about one in 300 of the neighborhood stars
have a mass in excess of 8 Msolar.
The main sequence ends at about 2100 K. Below that temperature are larger brown dwarfs: "failed"
stars, which have insufficient mass to initiate fusion.
Straight diagonal lines parallel to the tangent at the center of the main sequence correspond roughly to lines of equal
radius. Our B0 main sequence star has a radius around 10 Rsolar. The spacing of these lines of equal radius is roughly
logarithmic.
The first thing to notice is that by far, most of the data points lie on the main sequence. Since we are looking at a population of stars
at a different time in each star's life, that tells us that stars spend most of their lives on the main sequence. As a consequence, the
(super)giant and white dwarf regions of the diagram must represent stars at various stages near the ends of their lives (assuming that
most of the data points represent stars and not protostars).
Another thing to notice is that the white dwarf region of the diagram is distributed over a relatively wide range of luminosities and
temperatures. This means that knowledge of temperature does not translate as easily into knowledge about luminosity as it does with
main sequence stars. At first glance one might be tempted to say the same thing about the giant and supergiant regions, but because
they represent definite paths in the life cycles of those stars, and the stars themselves are much larger and brighter, they do not
represent so much of a problem.
We can learn much about the populations of stars in a galaxy like ours by looking statistically at the stars we know best: those
closest to us. Using a recent survey (source), we have culled spectral class, apparent magnitude and parallax data for 3633 stars
near the Sun. This data is available in an Excel spreadsheet. Using the parallax data, you can compute each star's distance from
Earth as
file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Stellar%20Populations.html

3/5

1/3/2016

Introduction to Physical Astronomy - Stellar Populations

distancepc = 1000 / parallaxmas


From the relation
absolute magnitude = apparent magnitude + 5 - 5 * log10 distancepc
you can then compute each star's absolute magnitude. The absolute magnitude can be used to determine the luminosity from the
relation
absolute magnitude = 4.83 - 2.5 * log10 luminositysolar
as
luminosity = 10(4.83 - absolute magnitude) / 2.5
Assuming that luminosity is approximately equal to mass3.5, you can then compute each star's mass as
masssolar = luminositysolar(1/3.5)
Portfolio Exercise: Create a histogram of the spectral classes of this population of stars. Omit stars with spectral
classes other than O, B, A, F, G, K or M, and stars with a range of possible classes. Include stars of all subclasses in
the main class (ie., count M and M1 through M9.5 all as class M). Note the percentage of stars for which there is no
subclass; what does this say about the difficulty of stellar classification?
Create a histogram of the known luminosity classes of this population of stars. Omit stars with luminosity classes other
than I-V and D/SD (white dwarf and sub-dwarf, sometimes designated VII and VI, respectively). Note the
percentage of stars for which there is no luminosity class; what does this say about the difficulty of luminosity
classification?
Create a histogram of the masses of this population of stars. What is the average mass to luminosity ratio (in solar
units)?

Binaries and Stellar Clusters


Approximately 75% of all stars are members of binary systems; the vast majority of these systems are primordial: they were
formed as a binary pair, as opposed to dynamical capture of one star by another.
Stellar clusters are groups of stars which are gravitationally bound (possibly very loosely). Clusters can be
open, or galactic clusters: these are small associations of young stars (100 to 10,000), usually found in the galactic plane;
as the gas clouds which gave them birth are blown away by stellar winds and/or ionizing radiation, the resulting loss of mass
allows most to eventually disperse; all stars begin life as cluster members;
globular clusters: larger (10,000 to millions of stars), older, and usually found in the galactic halo, these contain little
interstellar gas and usually have very dense centers; the average star has orbited that center on the order of 10,000 times
since it was formed.
You need a Java-capable browser to be able to use the applets. If they do not work with your Windows system, download
the Java VM (Virtual Machine) for your version of Windows at the download section at java.sun.com.
Table of Contents
References
Index
2013, Kenneth R. Koehler. All Rights Reserved. This document may be freely reproduced provided that this copyright notice is
included.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Stellar%20Populations.html

4/5

1/3/2016

Introduction to Physical Astronomy - Stellar Populations

Please send comments or suggestions to the author.

file:///E:/NSEA/Introduction%20to%20Physical%20Astronomy%20-%20Stellar%20Populations.html

5/5

You might also like