Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 23

1.

SPECTRUM OF ELECTROMAGNETIF WAVE


The electromagnetic spectrum is the range of all possible frequencies
of electromagnetic radiation. The "electromagnetic spectrum" of an object is the
characteristic distribution of electromagnetic radiation emitted or absorbed by that particular
object.
The electromagnetic spectrum extends from low frequencies used for
modern radio communication to gamma radiation at the short-wavelength ( high – frequency)
end, thereby covering wavelengths from thousands of kilometres down to a fraction of the
size of anatom. It is for this reason that the electromagnetic spectrum is highly studied for
spectroscopic purposes to characterize matter. The limit for long wavelength is the size of
the universe itself, while it is thought that the short wavelength limit is in the vicinity of
the Planck length, although in principle the spectrum is infinite and continuous.
History
For most of history, light was the only known part of the electromagnetic spectrum.
The ancient Greeks recognized that light traveled in straight lines and studied some of the
properties of it, including reflection and refraction. Over the years the study of light
continued and during the 16th and 17th centuries there were conflicting theories which
regarded light as either a wave or a particle. It was first linked to electromagnetism in 1845
when Michael Faraday noticed that light responded to a magnetic field. The first discovery of
electromagnetic waves other than light came in 1800, when William Herschel discovered
infrared light. He was studying the temperature of different colours by moving a thermometer
through light split by a prism. He noticed that the hottest temperature was beyond red. He
theorized that there was 'light' that you could not see. The next year, Johann Ritter worked at
the other end of spectrum and noticed that there were 'chemical rays' that behaved similar to,
but were beyond, visible violet light rays. They were later renamed ultraviolet radiation.
During the 1860s James Maxwell was studying electromagnetic field and realized that they
traveled at around the speed of light. He developed four partial differentialequations to
explain this correlation. These equations predicted many frequencies of electromagnetic
waves traveling at the speed of light. Attempting to prove Maxwell's equations, in
1886 Heinrich Hertz built an apparatus to generate and detect radio waves. He was able to
observe that they traveled at the speed of light and could be both reflected and refracted. In a
later experiment he similarly produced and measured microwaves. These new waves paved
the way for inventions such as the wireless telegraph and the radio. In 1895 Wilhelm
Röntgen noticed a new type of radiation emitted during an experiment. He called these x-rays
and found they were able to travel through parts of the human body but were reflected by
denser matter such as bones. Before long many uses were found for them in the field of
medicine. The last portion of the electromagnetic spectrum was filled in with the discovery
of gamma rays. In 1900 Paul Villard was studying radioactivity. He first thought they were
particles similar to alpha and beta particles. However, in 1910 Ernest Rutherford measured
their wave lengths and found that they were electromagnetic waves.
Range of the spectrum
Electromagnetic waves are typically described by any of the following three physical
properties:the frequency f, wavelength λ ,or photonenergy E. Frequencies range
23
from 2.4×10 Hz (1 GeV gamma rays) down to the local plasma frequency of the ionized
interstellar medium (~1 kHz). Wavelength is inversely proportional to the wave frequency, so
gamma rays have very short wavelengths that are fractions of the size of atoms, whereas
wavelengths can be as long as the universe. Photon energy is directly proportional to the
wave frequency, so gamma rays have the highest energy (around a billion electron volts) and
radio waves have very low energy (around a femto electron volts). These relations are
illustrated by the following equations:

where:

 c = 299,792,458 m/s is the speed of light in vacuum and


 h = 6.62606896(33)×10−34 J s = 4.13566733(10)×10−15 eV s is Planck's constant.
Whenever electromagnetic waves exist in a medium with matter, their wavelength is
decreased. Wavelengths of electromagnetic radiation, no matter what medium they are
traveling through, are usually quoted in terms of the vacuum wavelength, although this is
not always explicitly stated.
Generally, EM radiation is classified by wavelength into radio
wave, microwave, terahertz (or sub-millimeter) radiation, infrared, the visible region we
perceive as light, ultraviolet, X-rays and gamma rays. The behavior of EM radiation
depends on its wavelength. When EM radiation interacts with single atoms and
molecules, its behaviour also depends on the amount of energy per quantum (photon) it
carries.
Spectroscopy can detect a much wider region of the EM spectrum than the visible range
of 400 nm to 700 nm. A common laboratory spectroscope can detect wavelengths from
2 nm to 2500 nm. Detailed information about the physical properties of objects, gases, or
even stars can be obtained from this type of device. Spectroscopes are widely used
in astrophysics. For example, manyhydrogen atoms emit a radio wave photon that has a
wavelength of 21.12 cm. Also, frequencies of 30 Hz and below can be produced by and
are important in the study of certain stellar nebulae and frequencies as high
as 2.9×1027 Hz have been detected from astrophysical sources.
Rationale
Electromagnetic radiation interacts with matter in different ways in different parts of the
spectrum. The types of interaction can be so different that it seems to be justified to refer
to different types of radiation. At the same time, there is a continuum containing all these
"different kinds" of electromagnetic radiation. Thus we refer to a spectrum, but divide it
up based on the different interactions with matter.

Region of the
Main interactions with matter
spectrum

Collective oscillation of charge carriers in bulk material (plasma


Radio oscillation). An example would be the oscillation of the electrons
in an antenna.
Microwave through
Plasma oscillation, molecular rotation
far infrared

Near infrared Molecular vibration, plasma oscillation (in metals only)

Molecular electron excitation (including pigment molecules


Visible
found in the human retina), plasma oscillations (in metals only)

Excitation of molecular and atomic valence electrons, including


Ultraviolet
ejection of the electrons (photoelectric effect)

Excitation and ejection of core atomic electrons, Compton


X-rays
scattering (for low atomic numbers)

Energetic ejection of core electrons in heavy elements, Compton


Gamma rays scattering (for all atomic numbers), excitation of atomic nuclei,
including dissociation of nuclei

Creation of particle-antiparticle pairs. At very high energies a


High-energy gamma
single photon can create a shower of high-energy particles and
rays
antiparticles upon interaction with matter.
Types of radiation

The electromagnetic spectrum

The types of electromagnetic radiation are broadly classified into the following classes:[2]

1. Gamma radiation
2. X-ray radiation
3. Ultraviolet radiation
4. Visible radiation
5. Infrared radiation
6. Microwave radiation
7. Radio waves
This classification goes in the increasing order of wavelength, which is characteristic of
the type of radiation. While, in general, the classification scheme is accurate, in reality
there is often some overlap between neighboring types of electromagnetic energy. For
example, SLF radio waves at 60 Hz may be received and studied by astronomers, or may
be ducted along wires as electric power, although the latter is, in the strict sense, not
electromagnetic radiation at all (see near and far field) The distinction between X-rays
and gamma rays is based on sources: gamma rays are the photons generated from nuclear
decay or other nuclear and subnuclear/particle process, whereas X-rays are generated
by electronic transitions involving highly energetic inner atomic electrons. In general,
nuclear transitions are much more energetic than electronic transitions, so gamma-rays
are more energetic than X-rays, but exceptions exist. By analogy to electronic
transitions,muonic atom transitions are also said to produce X-rays, even though their
energy may exceed 6 megaelectronvolts (0.96 pJ), whereas there are many (77 known to
be less than 10 keV (1.6 fJ)) low-energy nuclear transitions (e.g., the 7.6 eV (1.22 aJ)
nuclear transition of thorium-229), and, despite being one million-fold less energetic than
some muonic X-rays, the emitted photons are still called gamma rays due to their nuclear
origin.
Also, the region of the spectrum of the particular electromagnetic radiation is reference
frame-dependent (on account of the Doppler shift for light), so EM radiation that one
observer would say is in one region of the spectrum could appear to an observer moving
at a substantial fraction of the speed of light with respect to the first to be in another part
of the spectrum. For example, consider the cosmic microwave background. It was
produced, when matter and radiation decoupled, by the de-excitation of hydrogen atoms
to the ground state. These photons were fromLyman series transitions, putting them in the
ultraviolet (UV) part of the electromagnetic spectrum. Now this radiation has undergone
enough cosmological red shift to put it into the microwave region of the spectrum for
observers moving slowly (compared to the speed of light) with respect to the cosmos.
However, for particles moving near the speed of light, this radiation will be blue-
shifted in their rest frame. The highest-energy cosmic ray protons are moving such that,
in their rest frame, this radiation is blueshifted to high-energy gamma rays, which interact
with the proton to produce bound quark-antiquark pairs (pions). This is the source of
the GZK limit.
Radio frequency
Radio waves generally are utilized by antennas of appropriate size (according to the
principle of resonance), with wavelengths ranging from hundreds of meters to about one
millimeter. They are used for transmission of data, via modulation. Television, mobile
phones, wireless networking, and amateur radio all use radio waves. The use of the radio
spectrum is regulated by many governments throughfrequency allocation.
Radio waves can be made to carry information by varying a combination of the
amplitude, frequency, and phase of the wave within a frequency band. When EM
radiation impinges upon a conductor, it couples to the conductor, travels along it,
and induces an electric current on the surface of that conductor by exciting the electrons
of the conducting material. This effect (the skin effect) is used in antennas.
Microwaves

Plot of Earth's atmospheric transmittance (or opacity) to various wavelengths of


electromagnetic radiation.

The super-high frequency (SHF) and extremely high frequency (EHF)


of microwaves come after radio waves. Microwaves are waves that are typically short
enough to employ tubular metal waveguides of reasonable diameter. Microwave energy
is produced with klystron andmagnetron tubes, and with solid state diodes such
as Gunn and IMPATT devices. Microwaves are absorbed by molecules that have adipole
moment in liquids. In a microwave oven, this effect is used to heat food. Low-intensity
microwave radiation is used in Wi-Fi, although this is at intensity levels unable to cause
thermal heating.
Volumetric heating, as used by microwave ovens, transfers energy through the material
electromagnetically, not as a thermal heat flux. The benefit of this is a more uniform
heating and reduced heating time; microwaves can heat material in less than 1% of the
time of conventional heating methods.
When active, the average microwave oven is powerful enough to cause interference at
close range with poorly shielded electromagnetic fields such as those found in mobile
medical devices and cheap consumer electronics.
Terahertz radiation
Terahertz radiation is a region of the spectrum between far infrared and microwaves.
Until recently, the range was rarely studied and few sources existed for microwave
energy at the high end of the band (sub-millimetre waves or so-called terahertz waves),
but applications such as imaging and communications are now appearing. Scientists are
also looking to apply terahertz technology in the armed forces, where high-frequency
waves might be directed at enemy troops to incapacitate their electronic equipment.[12]
Infrared radiation
The infrared part of the electromagnetic spectrum covers the range from roughly
300 GHz (1 mm) to 400 THz (750 nm). It can be divided into three parts:[2]

 Far-infrared, from 300 GHz (1 mm) to 30 THz (10 μm). The lower part of this range
may also be called microwaves. This radiation is typically absorbed by so-called
rotational modes in gas-phase molecules, by molecular motions in liquids, and
by phonons in solids. The water in Earth's atmosphere absorbs so strongly in this
range that it renders the atmosphere in effect opaque. However, there are certain
wavelength ranges ("windows") within the opaque range that allow partial
transmission, and can be used for astronomy. The wavelength range from
approximately 200 μm up to a few mm is often referred to as "sub-millimetre" in
astronomy, reserving far infrared for wavelengths below 200 μm.
 Mid-infrared, from 30 to 120 THz (10 to 2.5 μm). Hot objects (black-body radiators)
can radiate strongly in this range. It is absorbed by molecular vibrations, where the
different atoms in a molecule vibrate around their equilibrium positions. This range is
sometimes called the fingerprint region, since the mid-infrared absorption spectrum
of a compound is very specific for that compound.
 Near-infrared, from 120 to 400 THz (2,500 to 750 nm). Physical processes that are
relevant for this range are similar to those for visible light.
Visible radiation (light)
Above infrared in frequency comes visible light. This is the range in which
the sun and other stars emit most of their radiation and the spectrum that the human eye is
the most sensitiveto. Visible light (and near-infrared light) is typically absorbed and
emitted by electrons in molecules and atoms that move from one energy level to another.
The light we see with our eyes is really a very small portion of the electromagnetic
spectrum. A rainbow shows the optical (visible) part of the electromagnetic spectrum;
infrared (if you could see it) would be located just beyond the red side of the rainbow
with ultraviolet appearing just beyond the violet end.
Electromagnetic radiation with a wavelength between 380 nm and 760 nm (790–400
terahertz) is detected by the human eye and perceived as visible light. Other wavelengths,
especially near infrared (longer than 760 nm) and ultraviolet (shorter than 380 nm) are
also sometimes referred to as light, especially when the visibility to humans is not
relevant. White light is a combination of lights of different wavelengths in the visible
spectrum. Passing white light through a prism splits it up in to the several colors of light
observed in the visible spectrum between 400 nm and 780 nm.
If radiation having a frequency in the visible region of the EM spectrum reflects off an
object, say, a bowl of fruit, and then strikes our eyes, this results in our visual
perception of the scene. Our brain's visual system processes the multitude of reflected
frequencies into different shades and hues, and through this not-entirely-understood
psychophysical phenomenon, most people perceive a bowl of fruit.
At most wavelengths, however, the information carried by electromagnetic radiation is
not directly detected by human senses. Natural sources produce EM radiation across the
spectrum, and our technology can also manipulate a broad range of wavelengths. Optical
fiber transmits light that, although not necessarily in the visible part of the spectrum, can
carry information. The modulation is similar to that used with radio waves.
Ultraviolet light

The amount of penetration of UV relative to altitude in Earth's ozone

Next in frequency comes ultraviolet (UV). The wavelength of UV rays is shorter than
the violet end of the visible spectrum but longer than the X-ray.
UV in the very shortest range (next to X-rays) is cabable even of ionizing atoms
(see photoelectric effect), greatly changing their physical behavior.
At the middle range of UV, UV rays cannot ionize but can break chemical bonds, making
molecules to be unusually reactive. Sunburn, for example, is caused by the disruptive
effects of middle range UV radiation on skin cells, which is the main cause of skin
cancer. UV rays in the middle range can irreparably damage the complex DNA molecules
in the cells producing thymine dimers making it a very potent mutagen.
The sun emits a large amount of UV radiation, which could potentially turn Earth's land
surface into a barren desert (although ocean water would provide some protection for life
there). However, most of the Sun's most-damaging UV wavelengths are absorbed by the
atmosphere's nitrogen, oxygen, and ozone layerbefore they reache the surface. The higher
ranges of UV (vacuum UV) are absorbed by nitrogen and (at longer wavelengths by
simple diatomic oxygen in the air. Most of the UV in this mid-range is blocked by the
ozone layer, which absorbs strongly in the important 200-315 nm range, the lower part of
which is too long to be absorbed by ordinary dioxygen (oxygen) in air. The range
between 315 nm and visibile light (called UV-A) is not blocked well by the atmosphere,
but does not cause sunburn and does less biological damage. However, it is not harmless
and does cause oxygen radicals, mutation and skin damage. Seeultraviolet for more
information.
X-rays
After UV come X-rays, which, like the upper ranges of UV are also ionizing. However,
due to their higher energies, X-rays can also interact with matter by means of
the Compton effect. Hard X-rays have shorter wavelengths than soft X-rays. As they can
pass through most substances, X-rays can be used to 'see through' objects, the most
notable use being diagnostic X-ray images in medicine (a process known as radiography),
as well as for high-energy physics and astronomy. Neutron stars and accretion disks
around black holes emit X-rays, which enable us to study them. X-rays are given off by
stars and are strongly emitted by some types of nebulae.
Gamma rays
After hard X-rays come gamma rays, which were discovered by Paul Villard in 1900.
These are the most energetic photons, having no defined lower limit to their wavelength.
They are useful toastronomers in the study of high-energy objects or regions, and find a
use with physicists thanks to their penetrative ability and their production
from radioisotopes. Gamma rays are also used for theirradiation of food and seed for
sterilization, and in medicine they are used in radiation cancer therapy and some kinds of
diagnostic imaging such as PET scans. The wavelength of gamma rays can be measured
with high accuracy by means of Compton scattering.
Note that there are no precisely defined boundaries between the bands of the
electromagnetic spectrum. Radiation of some types have a mixture of the properties of
those in two regions of the spectrum. For example, red light resembles infrared radiation
in that it can resonate some chemical bonds.

Regions of the Electromagnetic Spectrum


The following table gives approximate wavelengths, frequencies, and energies
for selected regions of the electromagnetic spectrum.

Spectrum of Electromagnetic Radiation

Region Wavelength Wavelength Frequency Energy


(Angstroms) (centimeters) (Hz) (eV)

Radio > 109 > 10 < 3 x 109 < 10-5

Microwave 109 - 106 10 - 0.01 3 x 109 - 3 x 1012 10-5 - 0.01

Infrared 106 - 7000 0.01 - 7 x 10-5 3 x 1012 - 4.3 x 1014 0.01 - 2

Visible 7000 - 4000 7 x 10-5 - 4 x 10-5 4.3 x 1014 - 7.5 x 1014 2-3

Ultraviolet 4000 - 10 4 x 10-5 - 10-7 7.5 x 1014 - 3 x 1017 3 - 103

X-Rays 10 - 0.1 10-7 - 10-9 3 x 1017 - 3 x 1019 103 - 105

Gamma Rays < 0.1 < 10-9 > 3 x 1019 > 105
2. MAGNITUDE OF STAR

Ancient Origins
Star magnitudes do count backward, the result of an ancient fluke that seemed like a
good idea at the time. The story begins around 129 B.C., when the Greek
astronomer Hipparchusproduced the first well-known star catalog. Hipparchus ranked his
stars in a simple way. He called the brightest ones "of the first magnitude," simply meaning
"the biggest." Stars not so bright he called "of the second magnitude," or second biggest. The
faintest stars he could see he called "of the sixth magnitude." Around A.D. 140 Claudius
Ptolemy copied this system in his own star list. Sometimes Ptolemy added the words
"greater" or "smaller" to distinguish between stars within a magnitude class. Ptolemy's works
remained the basic astronomy texts for the next 1,400 years, so everyone used the system of
first to sixth magnitudes. It worked just fine.

Galileo forced the first change. On turning his newly made telescopes to the sky, Galileo
discovered that stars existed that were fainter than Ptolemy's sixth magnitude. "Indeed, with
the glass you will detect below stars of the sixth magnitude such a crowd of others that
escape natural sight that it is hardly believable," he exulted in his 1610 tract Sidereus
Nuncius. "The largest of these . . . we may designate as of the seventh magnitude." Thus did a
new term enter the astronomical language, and the magnitude scale became open-ended.
There could be no turning back.

As telescopes got bigger and better, astronomers kept adding more magnitudes to the bottom
of the scale. Today a pair of 50-millimeter binoculars will show stars of about 9th magnitude,
a 6-inch amateur telescope will reach to 13th magnitude, and the Hubble Space Telescope has
seen objects as faint as 31st magnitude.

By the middle of the 19th century, astronomers realized there was a pressing need to define
the entire magnitude scale more precisely than by eyeball judgment. They had already
determined that a 1st-magnitude star shines with about 100 times the light of a 6th-magnitude
star. Accordingly, in 1856 the Oxford astronomer Norman R. Pogson proposed that a
difference of five magnitudes be exactly defined as a brightness ratio of 100 to 1. This
convenient rule was quickly adopted. One magnitude thus corresponds to a brightness
difference of exactly the fifth root of 100, or very close to 2.512 — a value known as the
Pogson ratio.

Brightness of stars is assigned a number starting with the brightest star starting at about -1
magnitude. Dimmer stars are zero or positive numbers. The larger the number means the
dimmer the star is. For example, a star -1 magnitude is brighter than a star 0 magnitude. A
star 0 magnitude is brighter than a star 1 magnitude. A star 1 magnitude is brighter than a star
2 magnitude. A star 4 magnitude is brighter than a star 5 magnitude. Magnitude sequence for
stars starting with the brightest is -1, 0, 1, 2, 3, 4, 5, 6, 7, 8, 9, 10 magnitude, ... etc.

The decimal point is not used when star magnitudes are used on a star map. The decimal
point could be confused for a star on the map. At the top of this page is the constellationUrsa
Minor with star magnitudes for some of its stars. For example, magnitude 31 on the star map
mean 3.1 and magnitude 55 on the star map mean 5.5.

Historicaly the magnitude system started with Hipparcus and Ptolemy when they divided the
stars into six magnitudes. About 20 of the brightest stars that they could observe from their
location were assigned to the first magnitude. The next set of bright stars were assigned to
second magnitude and so forth. Sixth magnitude stars were assigned to stars that were barely
visible to the unaided eye under favorable conditions. It was empirically determined that the
ratio of first magnitude to sixth magnitude was 100 to 1. A logarithmic scale of 2.512
between magnitude levels is implemented. For example, a first magnitude star is 100 brighter
than a sixth magnitude star or the sixth magnitude star is 1/100 or .01 dimmer that a first
magnitude star. Second example, a fifth magnitude star is 2.512 times brighter than a sixth
magnitude star or the sixth magnitude star is 1/2.512 or .40 dimmer that a fifth magnitude
star. A star is 2.512 times brighter than a star one magnitude less.

Six Star Magnitude Table

Logarithmic scale of
How Much Brighter
Star Magnitude 2.512 X between magnitude levels
than a Sixth Magnitude Star
Starting at Sixth Magnitude

1 100 Times 2.51 x 2.51 x 2.51 x 2.51 x 2.51

2 39.8 Times 2.51 x 2.51 x 2.51 x 2.51

3 15.8 Times 2.51 x 2.51 x 2.51

4 6.3 Times 2.51 x 2.51

5 2.51 Times 2.51 x

With the invention of the telescope and modern equipment to measure star magnitudes the
scale has been extended in both directions. Dimmer stars are assigned magnitudes larger than
6 ( 6, 7, 8, 9, ... 30th ... etc.) The Hubble Space Telescope Deep Field image contains some
galaxies as faint as 30th magnitude. First magnitude stars are corrected across the scale of 1,
0, -1 with the brightest star Sirius at -1.44. The scale increases in brightness with negative
numbers. For example, the brightest planet Venus varies in brightness and is about -4.4
magnitude at maximum brightness. The Moon is -12.7 magnitude at maximum brightness and
the Sun is -26.75 magnitude.

The below Star Magnitude Table Based on -1 Magnitude Star shows how much dimmer than
a -1 magnitude star are stars to 19th magnitude. For example, most 10 x 50 or 7 x 50
binoculars can detect a 9 magnitude star. A 9 magnitude star is one tenth thousand (1/10,000
or .0001) dimmer than -1 magnitude star.

Star Magnitude Table Showing How Much Dimmer


Other Magnitudes are as Compared to a -1 Magnitude Star

How Much Dimmer How Much Dimmer


Star Magnitude
than a -1 Magnitude Star than a -1 Magnitude Star

-1
0 1/2.51 0.398
1 1/6.31 0.158
2 1/15 0.063
3 1/39 0.0251
4 1/100 0.0100
5 1/251 0.00398
6 1/630 0.00158
7 1/1,584 0.000630
8 1/3,981 0.000251
9 1/10,000 0.000100
10 1/25,118 0.0000398
11 1/63,095 0.0000158
12 1/158,489 0.00000631
13 1/398,107 0.00000251
14 1/1,000,000 0.00000100
15 1/2,511,886 0.000000398
16 1/6,309,573 0.000000158
17 1/15,848,931 0.000000063
18 1/39,810,717 0.000000025
19 1/100,000,000 0.000000010

The stars of Ursa Minor is a good constellation to determine how faint of a star can be
observed. On star maps bright stars are represented with large dots while dimmer stars are
represented with smaller dots. The brightness of the stars of Ursa Minor get fainter starting
with Polaris at 2.0 magnitude which is located on the right side of the below star maps. The
rest of the stars starting from bright to dim are 2.1, 3.1, 4.2, 4.3, 4.4, 5.0, 5.2 and 5.5
magnitude. Also note that Polaris is located in the same place in the sky throughout the year
for each observing location. Because Portland, Oregon, U.S.A. is at north 45 degrees latitude
Polaris is always located half way between the directly overhead and due north.

The star magnitude scale measures the intensity of the brightness of stars.
The lower the number, the brighter the star, and the higher the magnitude, the
dimmer the star. A logarithmic scale is used because the brightness of any star is
2.5 times brighter than the star of 1 magnitude greater, so a star of the third
magnitude is 2.5 times brighter than a star of the fourth magnitude, and a star of
the second magnitude is 2.5 times brighter than a star of the third magnitude. A
logarithmic scale was chosen since the difference between the absolute magnitude
and the apparent magnitue depends on distance, but isn't much. If it were
exponential, it would still depend on distance, but the difference between the two
magnitudes would be very great. To find out how bright or dim a star is compared
to a star of the first magnitude, take 2.5x, x being 1 less than the magnitude of the
star. The equation is then:

2.5^x-1 = # of times dimmer than a star of the first magnitude.


Magnitude of star | times dimmer than 1st mag. star
1|0
2 | 2.5
3 | 6.25
4 | 15.63
5 | 39.06
6 | 97.66
7 | 244.14

To find out how many times brighter one star is than another, take 2.5^x, where x
is the number of levels of magnitude between the two stars. For instance, to find
out how much brighter a 3 star is to a 5 star, take 2.5^2, and you get 6.25, so a star
of the 3rd magnitude is 6.25 times brighter than a star of the 5th magnitude.

To find out the absolute magnitude when the apparent magnitude and distance in
parsecs is known (10 pc is 33 light years) the formula is:

M=m-5log(d/10).
3. DIAGRAM H-R
The Hertzsprung–Russell diagram is a scatter graph of stars showing the
relationship between the stars'absolute magnitudes or luminosities versus their spectral
types or classifications and effective temperatures. Hertzsprung–Russell diagrams
are not pictures or maps of the locations of the stars. Rather, they plot each star on a graph
measuring the star's absolute magnitude or brightness against its temperature and color.
Hertzsprung–Russell diagrams are also referred to by the abbreviation H–R
diagram or HRD. The diagram was created circa 1910 by Ejnar Hertzsprung and Henry
Norris Russell and represents a major step towards an understanding of stellar evolution or
"the lives of stars".
Forms of diagram
There are several forms of the Hertzsprung–Russell diagram, and the nomenclature is
not very well defined. The original diagram displayed the spectral type of stars on the
horizontal axis and the absolute magnitude on the vertical axis. The first quantity (i.e. spectral
type) is difficult to plot as it is not a numerical quantity and in modern versions of the chart it
is replaced by the B-V color index of the stars. This type of diagram is what is often called a
Hertzsprung–Russell diagram, or specifically a color-magnitude diagram (CMD), and it is
often used by observers. In cases where the stars are known to be at identical distances such
as with a star cluster, a color-magnitude diagram is often used to describe a plot of the stars in
the cluster in which the vertical axis is the apparent magnitude.
Another form of the diagram plots the effective surface temperature of the star on one axis
and the luminosity of the star on the other. This is what theoreticians calculate using
computer models that describe the evolution of stars. This type of diagram should probably
be called temperature-luminosity diagram, but this term is hardly ever used, the
term Hertzsprung–Russell diagram being preferred instead. One peculiar characteristic of this
form of the H–R diagram is that the temperatures are plotted from high temperature to low
temperature, which aids in comparing this form of the H–R diagram with the observational
form.
Although the two types of diagrams are similar, astronomers make a sharp distinction
between the two. The reason for this distinction is that the exact transformation from one to
the other is not trivial, and depends on the stellar-atmosphere model being used and its
parameters (like composition and pressure, apart from temperature and luminosity). Also, one
needs to know the distance to the observed objects and the degree of interstellar
reddening. Empirical transformations between various color indices and effective temperature
are available in literature.
The H–R diagram can be used to define different types of stars and to match theoretical
predictions of stellar evolution using computer models with observations of actual stars. It is
then necessary to convert either the calculated quantities to observables, or the other way
around, thus introducing an extra uncertainty.
Interpretation
Most of the stars occupy the region in the diagram along the line called the main sequence.
During that stage stars are fusing hydrogen in their cores. The next concentration of stars is
on the horizontal branch (helium fusion in the core and hydrogen burning in a shell
surrounding the core). Another prominent feature is the Hertzsprung gap located in the region
between A5 and G0 spectral type and between +1 and −3 absolute magnitudes (i.e. between
the top of the main sequence and the giants in the horizontal branch). RR Lyrae variable stars
can be found in the left of this gap. Cepheid variables reside in the upper section of
the instability strip.
The H-R diagram can also be used by scientists to roughly measure how far away a star
cluster is from Earth. This can be done by comparing the apparent magnitudes of the stars in
the cluster to the absolute magnitudes of stars with known distances (or of model stars). The
observed group is then shifted in the vertical direction, until the two main sequences overlap.
The difference in magnitude that was bridged in order to match the two groups is called
the distance modulus and is a direct measure for the distance. This technique is known
as main-sequence fitting or spectroscopic parallax.
Diagram's role in the development of stellar physics
Contemplation of the diagram led astronomers to speculate that it might
demonstrate stellar evolution, the main suggestion being that stars collapsed from red giants
to dwarf stars, then moving down along the line of the main sequence in the course of their
lifetimes. Stars were thought therefore to radiate energy by converting gravitational energy
into radiation through the Kelvin–Helmholtz mechanism. This mechanism resulted in an age
for the sun of only tens of millions of years, creating a conflict over the age of the solar
system between astronomers, and biologists and geologists who had evidence that the earth
was far older than that. This conflict was only resolved in the 1930s when nuclear fusion was
identified as the source of stellar energy.
However, following Russell's presentation of the diagram to a meeting of the Royal
Astronomical Society in 1912, Arthur Eddington was inspired to use it as a basis for
developing ideas on stellar physics. In 1926, in his book The Internal Constitution of the
Stars he explained the physics of how stars fit on the diagram. This was a particularly
remarkable development since at that time the major problem of stellar theory, the source of a
star's energy, was still unsolved. Thermonuclear energy, and even that stars are largely
composed of hydrogen (see metallicity), had yet to be discovered. Eddington managed to
sidestep this problem by concentrating on the thermodynamics of radiative transport of
energy in stellar interiors. So, Eddington predicted that dwarf stars remain in an essentially
static position on the main sequence for most of their lives. In the 1930s and 1940s, with an
understanding of hydrogen fusion, came a physically based theory of evolution to red giants,
and white dwarfs. By this time, study of the Hertzsprung–Russell diagram did not drive such
developments but merely allowed stellar evolution to be presented graphically.
Hertzsprung–Russell diagram[1] with 22,000 stars plotted from the Hipparcos catalog and
1,000 from the Gliese catalog of nearby stars. Stars tend to fall only into certain regions of
the diagram. The most predominant is the diagonal, going from the upper-left (hot and bright)
to the lower-right (cooler and less bright), called the main sequence. In the lower-left is
where white dwarfs are found, and above the main sequence are
the subgiants, giants and supergiants. The Sun is found on the main sequence at luminosity 1
(absolute magnitude 4.8) and B-V color index 0.66 (temperature 5780K, spectral type G2).
HR diagrams for two open clusters, M67 and NGC 188, showing the main sequence turn-off
at different ages.

Graphing or plotting data is an essential tool used by scientists. In attempting to make sense
of data and see if two quantities are related we can plot them and seek trends. If we have a
look at the two examples below the first shows two quantities, X and Y that an object may
have. When they are plotted we can see that there is no discernable relationship between X
and Y. In fact in this example there is no relationship, the data is purely random.
If we plot data for height versus mass for a small group of people, however, we see a very
different pattern as shown below.

As we might expect, there does appear to be a correlation between the height of a person and
their mass. In general, the taller a person is, the greater their mass but as with many other
characteristics of humans there is a large variation. Some people are tall and skinny, others
shorter but higher mass. There are, however, real physical limitations on both the height and
mass of people. We do not expect to find a 3.5 m person with a mass of 10 kg or a 1.0 m
person with a mass of 300 kg!

Credit: Dorritt Hoffleit, Yale University Credit: AIP Emilio Segre Visual
Observatory, courtesy AIP Emilio Segre Archives, Margaret Russell
Visual Archives Edmondson Collection

Ejnar Hertzsprung Henry Norris Russell

One of the most useful and powerful plots in astrophysics is the Hertzsprung-Russell diagram
(hereafter called the H-R diagram). It originated in 1911 when the Danish astronomer, Ejnar
Hertzsprung, plotted the absolute magnitude of stars against their colour (hence effective
temperature). Independently in 1913 the American astronomer Henry Norris Russell used
spectral class against absolute magnitude. Their resultant plots showed that the relationship
between temperature and luminosity of a star was not random but instead appeared to fall into
distinct groups. These are seen in the H-R diagram below. It has a few specific stars included
in the plot but otherwise just shows the main regions.

The majority of stars, including our Sun, are found along a region called the Main Sequence.
Main Sequence stars vary widely in effective temperature but the hotter they are, the more
luminous they are, hence the main sequence tends to follow a band going from the bottom
right of the diagram to the top left. These stars are fusing hydrogen to helium in their cores.
Stars spend the bulk of their existence as main sequence stars. Other major groups of stars
found on the H-R diagram are the giants and supergiants; luminous stars that have evolved
off the main sequence, and the white dwarfs. Whilst each of these types is discussed in detail
in later pages we can use their positions on the H-R diagram to infer some of their properties.

Using the H-R Diagram to Infer Stellar Properties


Let us look at the cool M-class stars as an example. If we look at the H-R diagram below we
can see that in fact there are three main groups of these stars.

At the bottom-right of the diagram we can see two named stars, Proxima Centauri and
Barnard's Star. These are both cool (approximately 2,500 K) and dim (absolute magnitudes of
about -13, only about 1/10,000 the luminosity of our Sun). Following the broad band straight
up we come across Mira, also cool but much more luminous. Travelling further up we come
across Antares and Betelgeuse. Again these stars are cool but they are extremely luminous,
almost 10,000× as luminous as the Sun. Why do these three groups differ so much in
luminosity?

The answer to this question depends upon the Stefan-Boltzmann relationship. You may recall
from equation 4.4 that the energy emitted per unit surface area per second is simply a
function of the fourth power of temperature, that is:

l ≈ σT4 (4.4)

If two stars have the same effective temperature they each have the same power output per
square metre of surface area. As the H-R diagram however shows that one is much more
luminous than the other it must have a greater total power output therefore must have a much
greater surface area - the more luminous star is bigger. We can see this from the full
expression for luminosity in equation 4.6:

L ≈ 4πR2σT4 (4.6)

The difference between the three groups of M-class stars is thus a difference in size. This is
acknowledged by the names given to each of the groups. The most luminous ones are
called supergiants (luminosity classes I and II), the luminous ones are
called giants (luminosity class III) and the dim ones are part of the main sequence
(luminosity class V) though historically the term dwarf stars was applied to this group.

If we look at the vertical band on the H-R diagram for hotter stars around type A spectral
class we see a similar pattern:

In this case the supergiants Rigel and Deneb have the same effective temperature as Sirius
but have extremely high luminosities. They have large radii than Sirius hence greater surface
areas and higher luminosities. Sirius is a main sequence star but because it is hotter than the
red main sequence Barnard's Star it is much more luminous than it. If you follow the pink
band for hot stars down to the bottom of the H-R diagram you will notice that it intersects
another group of stars that includes Procyon B. These are the white dwarfs. They are very hot
(about 10,000 K or hotter) therefore emit a lot of energy per second for each square metre of
their surface. The fact that they are so dim however, means that they must be extremely small
and have a very low surface area. The terminology of white dwarf must not be confused with
the old-fashioned term of dwarf stars that was applied to main sequence stars. White dwarfs
are very different objects to main sequence stars as we shall see in a later page. Technically
they have a luminosity class of wd. Simple calculations provide a size for white dwarfs
roughly that of our Earth, less than 1/100 that of the Sun.

If we compare the dimmest stars on the H-R diagram we can also make some inferences. The
following diagram shows the lower region of the H-R diagram.

Procyon B and Barnard's Star share the same low luminosity with an absolute magnitude of
about +13. Procyon B however is much hotter than Barnard's Star thus emits much more
energy per second per unit surface area. Given that they have the same total power output
Procyon B must therefore have less surface area than Barnard's Star, that is its radius is
smaller.

Axes on the H-R Diagram

This points to an interesting and sometimes confusing feature of the H-R diagram - the scales
on the axes. Unlike the height/mass plot earlier in this section, the effective temperature does
not increase as it goes from left to right, it actually decreases, that is the highest temperature
is on the left-hand side. If colour index (B-V) rather than effective temperature is used then it
goes from negative (blue) on the left to positive (red) on the right. A third alternative along
the horizontal axis is to use spectral class. Of course, all three quantities are essentially
showing the same thing. The diagram below shows the possible axes for an H-R diagram.
The vertical axis displays the luminosity of the stars. This is either as a ratio compared with
that of the Sun or as absolute magnitude, M. One point to be careful of when using absolute
magnitude is to remember that the lower or more negative the absolute magnitude, the more
luminous the star. The brightest stars therefore appear at the top of the H-R diagram with the
vertical axis having the most negative value of M at the top.

In some circumstances, such as when plotting stars in a specific open or globular cluster,
apparent magnitude, m, or V, rather than absolute magnitude may be used. This is valid as all
the stars in the cluster are effectively at the same distance away from us hence any
differences in apparent magnitude are due to actual difference in luminosity or M. Diagrams
where V is plotted against colour index, B-V, are also known as colour-magnitude diagrams.

You might also like