Astronomical spectroscopy

Last updated
The Star-Spectroscope of the Lick Observatory in 1898. Designed by James Keeler and constructed by John Brashear. Star-Spectroscope.jpg
The Star-Spectroscope of the Lick Observatory in 1898. Designed by James Keeler and constructed by John Brashear.

Astronomical spectroscopy is the study of astronomy using the techniques of spectroscopy to measure the spectrum of electromagnetic radiation, including visible light, ultraviolet, X-ray, infrared and radio waves that radiate from stars and other celestial objects. A stellar spectrum can reveal many properties of stars, such as their chemical composition, temperature, density, mass, distance and luminosity. Spectroscopy can show the velocity of motion towards or away from the observer by measuring the Doppler shift. Spectroscopy is also used to study the physical properties of many other types of celestial objects such as planets, nebulae, galaxies, and active galactic nuclei.

Contents

Background

Opacity of the Earth's atmosphere for different wavelengths of electromagnetic radiation. The atmosphere blocks some wavelengths but it is mostly transparent for visible light and a wide range of radio waves. Atmospheric electromagnetic opacity.svg
Opacity of the Earth's atmosphere for different wavelengths of electromagnetic radiation. The atmosphere blocks some wavelengths but it is mostly transparent for visible light and a wide range of radio waves.

Astronomical spectroscopy is used to measure three major bands of radiation in the electromagnetic spectrum: visible light, radio waves, and X-rays. While all spectroscopy looks at specific bands of the spectrum, different methods are required to acquire the signal depending on the frequency. Ozone (O3) and molecular oxygen (O2) absorb light with wavelengths under 300 nm, meaning that X-ray and ultraviolet spectroscopy require the use of a satellite telescope or rocket mounted detectors. [1] :27 Radio signals have much longer wavelengths than optical signals, and require the use of antennas or radio dishes. Infrared light is absorbed by atmospheric water and carbon dioxide, so while the equipment is similar to that used in optical spectroscopy, satellites are required to record much of the infrared spectrum. [2]

Optical spectroscopy

With a reflection grating, incident light is separated into several diffraction orders which separate different wavelengths apart (red and blue lines), excepting the 0-th order (black). Blazedgrating.jpg
With a reflection grating, incident light is separated into several diffraction orders which separate different wavelengths apart (red and blue lines), excepting the 0-th order (black).

Physicists have been looking at the solar spectrum since Isaac Newton first used a simple prism to observe the refractive properties of light. [3] In the early 1800s Joseph von Fraunhofer used his skills as a glassmaker to create very pure prisms, which allowed him to observe 574 dark lines in a seemingly continuous spectrum. [4] Soon after this, he combined telescope and prism to observe the spectrum of Venus, the Moon, Mars, and various stars such as Betelgeuse; his company continued to manufacture and sell high-quality refracting telescopes based on his original designs until its closure in 1884. [5] :28–29

The resolution of a prism is limited by its size; a larger prism will provide a more detailed spectrum, but the increase in mass makes it unsuitable for highly detailed work. [6] This issue was resolved in the early 1900s with the development of high-quality reflection gratings by J.S. Plaskett at the Dominion Observatory in Ottawa, Canada. [5] :11 Light striking a mirror will reflect at the same angle, however a small portion of the light will be refracted at a different angle; this is dependent upon the indices of refraction of the materials and the wavelength of the light. [7] By creating a "blazed" grating which utilizes a large number of parallel mirrors, the small portion of light can be focused and visualized. These new spectroscopes were more detailed than a prism, required less light, and could be focused on a specific region of the spectrum by tilting the grating. [6]

The limitation to a blazed grating is the width of the mirrors, which can only be ground a finite amount before focus is lost; the maximum is around 1000 lines/mm. In order to overcome this limitation holographic gratings were developed. Volume phase holographic gratings use a thin film of dichromated gelatin on a glass surface, which is subsequently exposed to a wave pattern created by an interferometer. This wave pattern sets up a reflection pattern similar to the blazed gratings but utilizing Bragg diffraction, a process where the angle of reflection is dependent on the arrangement of the atoms in the gelatin. The holographic gratings can have up to 6000 lines/mm and can be up to twice as efficient in collecting light as blazed gratings. Because they are sealed between two sheets of glass, the holographic gratings are very versatile, potentially lasting decades before needing replacement. [8]

Light dispersed by the grating or prism in a spectrograph can be recorded by a detector. Historically, photographic plates were widely used to record spectra until electronic detectors were developed, and today optical spectrographs most often employ charge-coupled devices (CCDs). The wavelength scale of a spectrum can be calibrated by observing the spectrum of emission lines of known wavelength from a gas-discharge lamp. The flux scale of a spectrum can be calibrated as a function of wavelength by comparison with an observation of a standard star with corrections for atmospheric absorption of light; this is known as spectrophotometry. [9]

Radio spectroscopy

Radio astronomy was founded with the work of Karl Jansky in the early 1930s, while working for Bell Labs. He built a radio antenna to look at potential sources of interference for transatlantic radio transmissions. One of the sources of noise discovered came not from Earth, but from the center of the Milky Way, in the constellation Sagittarius. [10] In 1942, JS Hey captured the Sun's radio frequency using military radar receivers. [1] :26 Radio spectroscopy started with the discovery of the 21-centimeter H I line in 1951.

Radio interferometry

Radio interferometry was pioneered in 1946, when Joseph Lade Pawsey, Ruby Payne-Scott and Lindsay McCready used a single antenna atop a sea cliff to observe 200 MHz solar radiation. Two incident beams, one directly from the sun and the other reflected from the sea surface, generated the necessary interference. [11] The first multi-receiver interferometer was built in the same year by Martin Ryle and Vonberg. [12] [13] In 1960, Ryle and Antony Hewish published the technique of aperture synthesis to analyze interferometer data. [14] The aperture synthesis process, which involves autocorrelating and discrete Fourier transforming the incoming signal, recovers both the spatial and frequency variation in flux. [15] The result is a 3D image whose third axis is frequency. For this work, Ryle and Hewish were jointly awarded the 1974 Nobel Prize in Physics. [16]

X-ray spectroscopy

Stars and their properties

Spectral-lines-continuous.svg
Continuous spectrum
Spectral-lines-absorption.svg
Absorption lines (continuum plus discrete spectrum)

Chemical properties

Newton used a prism to split white light into a spectrum of color, and Fraunhofer's high-quality prisms allowed scientists to see dark lines of an unknown origin. In the 1850s, Gustav Kirchhoff and Robert Bunsen described the phenomena behind these dark lines. Hot solid objects produce light with a continuous spectrum, hot gases emit light at specific wavelengths, and hot solid objects surrounded by cooler gases show a near-continuous spectrum with dark lines corresponding to the emission lines of the gases. [5] :42–44 [17] By comparing the absorption lines of the Sun with emission spectra of known gases, the chemical composition of stars can be determined.

The major Fraunhofer lines, and the elements with which they are associated, appear in the following table. Designations from the early Balmer Series are shown in parentheses.

DesignationElementWavelength (nm)
y O2 898.765
ZO2822.696
AO2759.370
BO2686.719
C (Hα) H 656.281
aO2627.661
D1 Na 589.592
D2Na588.995
D3 or d He 587.5618
e Hg 546.073
E2 Fe 527.039
b1 Mg 518.362
b2Mg517.270
b3Fe516.891
b4Mg516.733
DesignationElementWavelength (nm)
cFe495.761
F (Hβ)H486.134
dFe466.814
eFe438.355
G' (Hγ)H434.047
GFe430.790
G Ca 430.774
h (Hδ)H410.175
HCa+396.847
KCa+393.368
LFe382.044
NFe358.121
P Ti +336.112
TFe302.108
t Ni 299.444

Not all of the elements in the Sun were immediately identified. Two examples are listed below:

To date more than 20 000 absorption lines have been listed for the Sun between 293.5 and 877.0 nm, yet only approximately 75% of these lines have been linked to elemental absorption. [1] :69

By analyzing the equivalent width of each spectral line in an emission spectrum, both the elements present in a star and their relative abundances can be determined. [7] Using this information stars can be categorized into stellar populations; Population I stars are the youngest stars and have the highest metal content (the Sun is a Pop I star), while Population III stars are the oldest stars with a very low metal content. [19] [20]

Temperature and size

Black body curves for various temperatures. Black body.svg
Black body curves for various temperatures.

In 1860 Gustav Kirchhoff proposed the idea of a black body, a material that emits electromagnetic radiation at all wavelengths. [21] [22] In 1894 Wilhelm Wien derived an expression relating the temperature (T) of a black body to its peak emission wavelength (λmax): [23]

b is a constant of proportionality called Wien's displacement constant, equal to 2.897771955...×10−3 m⋅K. [24] This equation is called Wien's Law. By measuring the peak wavelength of a star, the surface temperature can be determined. [17] For example, if the peak wavelength of a star is 502 nm the corresponding temperature will be 5772 kelvins.

The luminosity of a star is a measure of the electromagnetic energy output in a given amount of time. [25] Luminosity (L) can be related to the temperature (T) of a star by:

,

where R is the radius of the star and σ is the Stefan–Boltzmann constant, with a value of 5.670374419...×10−8 W⋅m−2⋅K−4. [26] Thus, when both luminosity and temperature are known (via direct measurement and calculation) the radius of a star can be determined.

Galaxies

The spectra of galaxies look similar to stellar spectra, as they consist of the combined light of billions of stars.

Doppler shift studies of galaxy clusters by Fritz Zwicky in 1937 found that the galaxies in a cluster were moving much faster than seemed to be possible from the mass of the cluster inferred from the visible light. Zwicky hypothesized that there must be a great deal of non-luminous matter in the galaxy clusters, which became known as dark matter. [27] Since his discovery, astronomers have determined that a large portion of galaxies (and most of the universe) is made up of dark matter. In 2003, however, four galaxies (NGC 821, NGC 3379, NGC 4494, and NGC 4697) were found to have little to no dark matter influencing the motion of the stars contained within them; the reason behind the lack of dark matter is unknown. [28]

In the 1950s, strong radio sources were found to be associated with very dim, very red objects. When the first spectrum of one of these objects was taken there were absorption lines at wavelengths where none were expected. It was soon realised that what was observed was a normal galactic spectrum, but highly red shifted. [29] [30] These were named quasi-stellar radio sources, or quasars, by Hong-Yee Chiu in 1964. [31] Quasars are now thought to be galaxies formed in the early years of our universe, with their extreme energy output powered by super-massive black holes. [30]

The properties of a galaxy can also be determined by analyzing the stars found within them. NGC 4550, a galaxy in the Virgo Cluster, has a large portion of its stars rotating in the opposite direction as the other portion. It is believed that the galaxy is the combination of two smaller galaxies that were rotating in opposite directions to each other. [32] Bright stars in galaxies can also help determine the distance to a galaxy, which may be a more accurate method than parallax or standard candles. [33]

Interstellar medium

The interstellar medium is matter that occupies the space between star systems in a galaxy. 99% of this matter is gaseous – hydrogen, helium, and smaller quantities of other ionized elements such as oxygen. The other 1% is dust particles, thought to be mainly graphite, silicates, and ices. [34] Clouds of the dust and gas are referred to as nebulae.

There are three main types of nebula: absorption, reflection, and emission nebulae. Absorption (or dark) nebulae are made of dust and gas in such quantities that they obscure the starlight behind them, making photometry difficult. Reflection nebulae, as their name suggest, reflect the light of nearby stars. Their spectra are the same as the stars surrounding them, though the light is bluer; shorter wavelengths scatter better than longer wavelengths. Emission nebulae emit light at specific wavelengths depending on their chemical composition. [34]

Gaseous emission nebulae

In the early years of astronomical spectroscopy, scientists were puzzled by the spectrum of gaseous nebulae. In 1864 William Huggins noticed that many nebulae showed only emission lines rather than a full spectrum like stars. From the work of Kirchhoff, he concluded that nebulae must contain "enormous masses of luminous gas or vapour." [35] However, there were several emission lines that could not be linked to any terrestrial element, brightest among them lines at 495.9 nm and 500.7 nm. [36] These lines were attributed to a new element, nebulium, until Ira Bowen determined in 1927 that the emission lines were from highly ionised oxygen (O+2). [37] [38] These emission lines could not be replicated in a laboratory because they are forbidden lines; the low density of a nebula (one atom per cubic centimetre) [34] allows for metastable ions to decay via forbidden line emission rather than collisions with other atoms. [36]

Not all emission nebulae are found around or near stars where solar heating causes ionisation. The majority of gaseous emission nebulae are formed of neutral hydrogen. In the ground state neutral hydrogen has two possible spin states: the electron has either the same spin or the opposite spin of the proton. When the atom transitions between these two states, it releases an emission or absorption line of 21 cm. [34] This line is within the radio range and allows for very precise measurements: [36]

Using this information, the shape of the Milky Way has been determined to be a spiral galaxy, though the exact number and position of the spiral arms is the subject of ongoing research. [39]

Complex molecules

Dust and molecules in the interstellar medium not only obscures photometry, but also causes absorption lines in spectroscopy. Their spectral features are generated by transitions of component electrons between different energy levels, or by rotational or vibrational spectra. Detection usually occurs in radio, microwave, or infrared portions of the spectrum. [40] The chemical reactions that form these molecules can happen in cold, diffuse clouds [41] or in dense regions illuminated with ultraviolet light. [42] Most known compounds in space are organic, ranging from small molecules e.g. acetylene C2H2 and acetone (CH3)2CO; [43] to entire classes of large molecule e.g. fullerenes [42] and polycyclic aromatic hydrocarbons; to solids, such as graphite or other sooty material. [44]

Motion in the universe

Redshift and blueshift Redshift blueshift.svg
Redshift and blueshift

Stars and interstellar gas are bound by gravity to form galaxies, and groups of galaxies can be bound by gravity in galaxy clusters. [45] With the exception of stars in the Milky Way and the galaxies in the Local Group, almost all galaxies are moving away from Earth due to the expansion of the universe. [18]

Doppler effect and redshift

The motion of stellar objects can be determined by looking at their spectrum. Because of the Doppler effect, objects moving towards someone are blueshifted, and objects moving away are redshifted. The wavelength of redshifted light is longer, appearing redder than the source. Conversely, the wavelength of blueshifted light is shorter, appearing bluer than the source light:

where is the emitted wavelength, is the velocity of the object, and is the observed wavelength. Note that v<0 corresponds to λ<λ0, a blueshifted wavelength. A redshifted absorption or emission line will appear more towards the red end of the spectrum than a stationary line. In 1913 Vesto Slipher determined the Andromeda Galaxy was blueshifted, meaning it was moving towards the Milky Way. He recorded the spectra of 20 other galaxies — all but four of which were redshifted — and was able to calculate their velocities relative to the Earth. Edwin Hubble would later use this information, as well as his own observations, to define Hubble's law: The further a galaxy is from the Earth, the faster it is moving away. [18] [46] Hubble's law can be generalised to:

where is the velocity (or Hubble Flow), is the Hubble Constant, and is the distance from Earth.

Redshift (z) can be expressed by the following equations: [47]

Calculation of redshift,
Based on wavelengthBased on frequency

In these equations, frequency is denoted by and wavelength by . The larger the value of z, the more redshifted the light and the farther away the object is from the Earth. As of January 2013, the largest galaxy redshift of z~12 was found using the Hubble Ultra-Deep Field, corresponding to an age of over 13 billion years (the universe is approximately 13.82 billion years old). [48] [49] [50]

The Doppler effect and Hubble's law can be combined to form the equation , where c is the speed of light.

Peculiar motion

Objects that are gravitationally bound will rotate around a common center of mass. For stellar bodies, this motion is known as peculiar velocity and can alter the Hubble Flow. Thus, an extra term for the peculiar motion needs to be added to Hubble's law: [51]

This motion can cause confusion when looking at a solar or galactic spectrum, because the expected redshift based on the simple Hubble law will be obscured by the peculiar motion. For example, the shape and size of the Virgo Cluster has been a matter of great scientific scrutiny due to the very large peculiar velocities of the galaxies in the cluster. [52]

Binary stars

Two stars of different size orbiting the center of mass. The spectrum can be seen to split depending on the position and velocity of the stars. Wiki Spect Binaries v2.gif
Two stars of different size orbiting the center of mass. The spectrum can be seen to split depending on the position and velocity of the stars.

Just as planets can be gravitationally bound to stars, pairs of stars can orbit each other. Some binary stars are visual binaries, meaning they can be observed orbiting each other through a telescope. Some binary stars, however, are too close together to be resolved. [53] These two stars, when viewed through a spectrometer, will show a composite spectrum: the spectrum of each star will be added together. This composite spectrum becomes easier to detect when the stars are of similar luminosity and of different spectral class. [54]

Spectroscopic binaries can be also detected due to their radial velocity; as they orbit around each other one star may be moving towards the Earth whilst the other moves away, causing a Doppler shift in the composite spectrum. The orbital plane of the system determines the magnitude of the observed shift: if the observer is looking perpendicular to the orbital plane there will be no observed radial velocity. [53] [54] For example, a person looking at a carousel from the side will see the animals moving toward and away from them, whereas if they look from directly above they will only be moving in the horizontal plane.

Planets, asteroids, and comets

Planets, asteroids, and comets all reflect light from their parent stars and emit their own light. For cooler objects, including Solar System planets and asteroids, most of the emission is at infrared wavelengths we cannot see, but that are routinely measured with spectrometers. For objects surrounded by gas, such as comets and planets with atmospheres, further emission and absorption happens at specific wavelengths in the gas, imprinting the spectrum of the gas on that of the solid object. In the case of worlds with thick atmospheres or complete cloud cover (such as the gas giants, Venus, and Saturn's satellite Titan), the spectrum is mostly or completely due to the atmosphere alone. [55]

Planets

The reflected light of a planet contains absorption bands due to minerals in the rocks present for rocky bodies, or due to the elements and molecules present in the atmosphere. To date over 3,500 exoplanets have been discovered. These include so-called Hot Jupiters, as well as Earth-like planets. Using spectroscopy, compounds such as alkali metals, water vapor, carbon monoxide, carbon dioxide, and methane have all been discovered. [56]

Asteroids

Asteroids can be classified into three major types according to their spectra. The original categories were created by Clark R. Chapman, David Morrison, and Ben Zellner in 1975, and further expanded by David J. Tholen in 1984. In what is now known as the Tholen classification, the C-types are made of carbonaceous material, S-types consist mainly of silicates, and X-types are 'metallic'. There are other classifications for unusual asteroids. C- and S-type asteroids are the most common asteroids. In 2002 the Tholen classification was further "evolved" into the SMASS classification, expanding the number of categories from 14 to 26 to account for more precise spectroscopic analysis of the asteroids. [57] [58]

Comets

Optical spectrum of Comet Hyakutake. Spectrum of Comet Hyakutake.gif
Optical spectrum of Comet Hyakutake.

The spectra of comets consist of a reflected solar spectrum from the dusty clouds surrounding the comet, as well as emission lines from gaseous atoms and molecules excited to fluorescence by sunlight and/or chemical reactions. For example, the chemical composition of Comet ISON [59] was determined by spectroscopy due to the prominent emission lines of cyanogen (CN), as well as two- and three-carbon atoms (C2 and C3). [60] Nearby comets can even be seen in X-ray as solar wind ions flying to the coma are neutralized. The cometary X-ray spectra therefore reflect the state of the solar wind rather than that of the comet. [61]

See also

Israeli stamp featuring astronomical spectroscopy, 1964 Israeli stamps 1964 - Sixteenth Independence Day, B.jpg
Israeli stamp featuring astronomical spectroscopy, 1964

Related Research Articles

<span class="mw-page-title-main">Nebula</span> Body of interstellar clouds

A nebula is a distinct luminescent part of interstellar medium, which can consist of ionized, neutral, or molecular hydrogen and also cosmic dust. Nebulae are often star-forming regions, such as in the "Pillars of Creation" in the Eagle Nebula. In these regions, the formations of gas, dust, and other materials "clump" together to form denser regions, which attract further matter and eventually become dense enough to form stars. The remaining material is then thought to form planets and other planetary system objects.

<span class="mw-page-title-main">Quasar</span> Active galactic nucleus containing a supermassive black hole

A quasar is an extremely luminous active galactic nucleus (AGN). It is sometimes known as a quasi-stellar object, abbreviated QSO. The emission from an AGN is powered by a supermassive black hole with a mass ranging from millions to tens of billions of solar masses, surrounded by a gaseous accretion disc. Gas in the disc falling towards the black hole heats up and releases energy in the form of electromagnetic radiation. The radiant energy of quasars is enormous; the most powerful quasars have luminosities thousands of times greater than that of a galaxy such as the Milky Way. Quasars are usually categorized as a subclass of the more general category of AGN. The redshifts of quasars are of cosmological origin.

<span class="mw-page-title-main">Redshift</span> Change of wavelength in photons during travel

In physics, a redshift is an increase in the wavelength, and corresponding decrease in the frequency and photon energy, of electromagnetic radiation. The opposite change, a decrease in wavelength and simultaneous increase in frequency and energy, is known as a negative redshift, or blueshift. The terms derive from the colours red and blue which form the extremes of the visible light spectrum. The main causes of electromagnetic redshift in astronomy and cosmology are the relative motions of radiation sources, which give rise to the relativistic Doppler effect, and gravitational potentials, which gravitationally redshift escaping radiation. All sufficiently distant light sources show cosmological redshift corresponding to recession speeds proportional to their distances from Earth, a fact known as Hubble's law that implies the universe is expanding.

<span class="mw-page-title-main">Spectroscopy</span> Study involving matter and electromagnetic radiation

Spectroscopy is the field of study that measures and interprets the electromagnetic spectra that result from the interaction between electromagnetic radiation and matter as a function of the wavelength or frequency of the radiation. In simpler terms, spectroscopy is the precise study of color as generalized from visible light to all bands of the electromagnetic spectrum.

<span class="mw-page-title-main">Optical spectrometer</span> Instrument to measure the properties of visible light

An optical spectrometer is an instrument used to measure properties of light over a specific portion of the electromagnetic spectrum, typically used in spectroscopic analysis to identify materials. The variable measured is most often the irradiance of the light but could also, for instance, be the polarization state. The independent variable is usually the wavelength of the light or a unit directly proportional to the photon energy, such as reciprocal centimeters or electron volts, which has a reciprocal relationship to wavelength.

<span class="mw-page-title-main">Planetary nebula</span> Type of emission nebula created by dying red giants

A planetary nebula is a type of emission nebula consisting of an expanding, glowing shell of ionized gas ejected from red giant stars late in their lives.

<span class="mw-page-title-main">Interstellar medium</span> Matter and radiation in the space between the star systems in a galaxy

In astronomy, the interstellar medium (ISM) is the matter and radiation that exist in the space between the star systems in a galaxy. This matter includes gas in ionic, atomic, and molecular form, as well as dust and cosmic rays. It fills interstellar space and blends smoothly into the surrounding intergalactic space. The energy that occupies the same volume, in the form of electromagnetic radiation, is the interstellar radiation field. Although the density of atoms in the ISM is usually far below that in the best laboratory vacuums, the mean free path between collisions is short compared to typical interstellar lengths, so on these scales the ISM behaves as a gas (more precisely, as a plasma: it is everywhere at least slightly ionized), responding to pressure forces, and not as a collection of non-interacting particles.

<span class="mw-page-title-main">Seyfert galaxy</span> Class of active galaxies with very bright nuclei

Seyfert galaxies are one of the two largest groups of active galaxies, along with quasars. They have quasar-like nuclei with very high surface brightnesses whose spectra reveal strong, high-ionisation emission lines, but unlike quasars, their host galaxies are clearly detectable.

<span class="mw-page-title-main">H II region</span> Large, low-density interstellar cloud of partially ionized gas

An H II region or HII region is a region of interstellar atomic hydrogen that is ionized. It is typically in a molecular cloud of partially ionized gas in which star formation has recently taken place, with a size ranging from one to hundreds of light years, and density from a few to about a million particles per cubic centimetre. The Orion Nebula, now known to be an H II region, was observed in 1610 by Nicolas-Claude Fabri de Peiresc by telescope, the first such object discovered.

<span class="mw-page-title-main">Emission spectrum</span> Frequencies of light emitted by atoms or chemical compounds

The emission spectrum of a chemical element or chemical compound is the spectrum of frequencies of electromagnetic radiation emitted due to electrons making a transition from a high energy state to a lower energy state. The photon energy of the emitted photons is equal to the energy difference between the two states. There are many possible electron transitions for each atom, and each transition has a specific energy difference. This collection of different transitions, leading to different radiated wavelengths, make up an emission spectrum. Each element's emission spectrum is unique. Therefore, spectroscopy can be used to identify elements in matter of unknown composition. Similarly, the emission spectra of molecules can be used in chemical analysis of substances.

<span class="mw-page-title-main">Astrophysics</span> Subfield of astronomy

Astrophysics is a science that employs the methods and principles of physics and chemistry in the study of astronomical objects and phenomena. As one of the founders of the discipline, James Keeler, said, Astrophysics "seeks to ascertain the nature of the heavenly bodies, rather than their positions or motions in space–what they are, rather than where they are." Among the subjects studied are the Sun, other stars, galaxies, extrasolar planets, the interstellar medium and the cosmic microwave background. Emissions from these objects are examined across all parts of the electromagnetic spectrum, and the properties examined include luminosity, density, temperature, and chemical composition. Because astrophysics is a very broad subject, astrophysicists apply concepts and methods from many disciplines of physics, including classical mechanics, electromagnetism, statistical mechanics, thermodynamics, quantum mechanics, relativity, nuclear and particle physics, and atomic and molecular physics.

<span class="mw-page-title-main">Absorption spectroscopy</span> Spectroscopic techniques that measure the absorption of radiation

Absorption spectroscopy is spectroscopy that involves techniques that measure the absorption of electromagnetic radiation, as a function of frequency or wavelength, due to its interaction with a sample. The sample absorbs energy, i.e., photons, from the radiating field. The intensity of the absorption varies as a function of frequency, and this variation is the absorption spectrum. Absorption spectroscopy is performed across the electromagnetic spectrum.

The Balmer series, or Balmer lines in atomic physics, is one of a set of six named series describing the spectral line emissions of the hydrogen atom. The Balmer series is calculated using the Balmer formula, an empirical equation discovered by Johann Balmer in 1885.

<span class="mw-page-title-main">Sombrero Galaxy</span> Galaxy in the constellation Virgo

The Sombrero Galaxy is a peculiar galaxy of unclear classification in the constellation borders of Virgo and Corvus, being about 9.55 megaparsecs from the Milky Way galaxy. It is a member of the Virgo II Groups, a series of galaxies and galaxy clusters strung out from the southern edge of the Virgo Supercluster. It has an isophotal diameter of approximately 29.09 to 32.32 kiloparsecs, making it slightly bigger in size than the Milky Way.

<span class="mw-page-title-main">Extinction (astronomy)</span> Interstellar absorption and scattering of light

In astronomy, extinction is the absorption and scattering of electromagnetic radiation by dust and gas between an emitting astronomical object and the observer. Interstellar extinction was first documented as such in 1930 by Robert Julius Trumpler. However, its effects had been noted in 1847 by Friedrich Georg Wilhelm von Struve, and its effect on the colors of stars had been observed by a number of individuals who did not connect it with the general presence of galactic dust. For stars lying near the plane of the Milky Way which are within a few thousand parsecs of the Earth, extinction in the visual band of frequencies is roughly 1.8 magnitudes per kiloparsec.

<span class="mw-page-title-main">Diffuse interstellar bands</span>

Diffuse interstellar bands (DIBs) are absorption features seen in the spectra of astronomical objects in the Milky Way and other galaxies. They are caused by the absorption of light by the interstellar medium. Circa 500 bands have now been seen, in ultraviolet, visible and infrared wavelengths.

<span class="mw-page-title-main">Cosmic Origins Spectrograph</span>

The Cosmic Origins Spectrograph (COS) is a science instrument that was installed on the Hubble Space Telescope during Servicing Mission 4 (STS-125) in May 2009. It is designed for ultraviolet (90–320 nm) spectroscopy of faint point sources with a resolving power of ≈1,550–24,000. Science goals include the study of the origins of large scale structure in the universe, the formation and evolution of galaxies, and the origin of stellar and planetary systems and the cold interstellar medium. COS was developed and built by the Center for Astrophysics and Space Astronomy (CASA-ARL) at the University of Colorado at Boulder and the Ball Aerospace and Technologies Corporation in Boulder, Colorado.

<span class="mw-page-title-main">History of spectroscopy</span>

Modern spectroscopy in the Western world started in the 17th century. New designs in optics, specifically prisms, enabled systematic observations of the solar spectrum. Isaac Newton first applied the word spectrum to describe the rainbow of colors that combine to form white light. During the early 1800s, Joseph von Fraunhofer conducted experiments with dispersive spectrometers that enabled spectroscopy to become a more precise and quantitative scientific technique. Since then, spectroscopy has played and continues to play a significant role in chemistry, physics and astronomy. Fraunhofer observed and measured dark lines in the Sun's spectrum, which now bear his name although several of them were observed earlier by Wollaston.

<span class="mw-page-title-main">Spectrometer</span> Used to measure spectral components of light

A spectrometer is a scientific instrument used to separate and measure spectral components of a physical phenomenon. Spectrometer is a broad term often used to describe instruments that measure a continuous variable of a phenomenon where the spectral components are somehow mixed. In visible light a spectrometer can separate white light and measure individual narrow bands of color, called a spectrum. A mass spectrometer measures the spectrum of the masses of the atoms or molecules present in a gas. The first spectrometers were used to split light into an array of separate colors. Spectrometers were developed in early studies of physics, astronomy, and chemistry. The capability of spectroscopy to determine chemical composition drove its advancement and continues to be one of its primary uses. Spectrometers are used in astronomy to analyze the chemical composition of stars and planets, and spectrometers gather data on the origin of the universe.

<span class="mw-page-title-main">TON 618</span> Quasar and Lyman-alpha blob in the constellation Canes Venatici

TON 618 is a hyperluminous, broad-absorption-line, radio-loud quasar and Lyman-alpha blob located near the border of the constellations Canes Venatici and Coma Berenices, with the projected comoving distance of approximately 18.2 billion light-years from Earth. It possesses one of the most massive black holes ever found, at 40 billion M.

References

  1. 1 2 3 4 Foukal, Peter V. (2004). Solar Astrophysics. Weinheim: Wiley VCH. p. 69. ISBN   3-527-40374-4.
  2. "Cool Cosmos – Infrared Astronomy". California Institute of Technology. Archived from the original on 11 October 2018. Retrieved 23 October 2013.
  3. Newton, Isaac (1705). Opticks: Or, A Treatise of the Reflections, Refractions, Inflections and Colours of Light. London: Royal Society. pp. 13–19.
  4. Fraunhofer, Joseph (1817). "Bestimmung des Brechungs- und des Farben-Zerstreuungs – Vermögens verschiedener Glasarten, in Bezug auf die Vervollkommnung achromatischer Fernröhre". Annalen der Physik. 56 (7): 282–287. Bibcode:1817AnP....56..264F. doi:10.1002/andp.18170560706.
  5. 1 2 3 4 Hearnshaw, J.B. (1986). The analysis of starlight. Cambridge: Cambridge University Press. ISBN   0-521-39916-5.
  6. 1 2 Kitchin, C.R. (1995). Optical Astronomical Spectroscopy. Bristol: Institute of Physics Publishing. pp. 127, 143. ISBN   0-7503-0346-8.
  7. 1 2 Ball, David W. (2001). Basics of Spectroscopy. Bellingham, Washington: Society of Photo-Optical Instrumentation Engineers. pp. 24, 28. ISBN   0-8194-4104-X.
  8. Barden, S.C.; Arns, J.A.; Colburn, W.S. (July 1998). d'Odorico, Sandro (ed.). "Volume-phase holographic gratings and their potential for astronomical applications" (PDF). Proc. SPIE. Optical Astronomical Instrumentation. 3355: 866–876. Bibcode:1998SPIE.3355..866B. CiteSeerX   10.1.1.25.5736 . doi:10.1117/12.316806. S2CID   17445305. Archived from the original (PDF) on 2010-07-28. Retrieved 2019-09-12.
  9. Oke, J. B.; Gunn, J. E. (1983). "Secondary standard stars for absolute spectrophotometry". The Astrophysical Journal. 266: 713. Bibcode:1983ApJ...266..713O. doi:10.1086/160817.
  10. Ghigo, F. "Karl Jansky". National Radio Astronomy Observatory. Associated Universities, Inc. Retrieved 24 October 2013.
  11. Pawsey, Joseph; Payne-Scott, Ruby; McCready, Lindsay (1946). "Radio-Frequency Energy from the Sun". Nature . 157 (3980): 158–159. Bibcode:1946Natur.157..158P. doi:10.1038/157158a0. PMID   21015114. S2CID   4056021.
  12. Ryle, M.; Vonberg, D. D. (1946). "Solar Radiation on 175 Mc./s". Nature. 158 (4010): 339–340. Bibcode:1946Natur.158..339R. doi:10.1038/158339b0. S2CID   4097569.
  13. Robertson, Peter (1992). Beyond southern skies: radio astronomy and the Parkes telescope. University of Cambridge. pp. 42, 43. ISBN   0-521-41408-3.
  14. W. E. Howard. "A Chronological History of Radio Astronomy" (PDF). Archived from the original (PDF) on 2012-07-14. Retrieved 2 December 2013.
  15. "How Radio Telescopes Work". Archived from the original on 3 December 2013. Retrieved 2 December 2013.
  16. "Press Release: The 1974 Nobel Prize in Physics" . Retrieved 2 December 2013.
  17. 1 2 Jenkins, Francis A.; Harvey E. White (1957). Fundamentals of Optics (4th ed.). New York: McGraw-Hill. pp.  430–437. ISBN   0-07-085346-0.
  18. 1 2 3 Morison, Ian (2008). Introduction to Astronomy and Cosmology (PDF). Wiley-Blackwell. p. 61. ISBN   978-0-470-03333-3. Archived from the original (PDF) on 2013-10-29.
  19. Gregory, Stephen A.; Michael Zeilik (1998). Introductory astronomy & astrophysics (4. ed.). Fort Worth [u.a.]: Saunders College Publ. p. 322. ISBN   0-03-006228-4.
  20. Pan, Liubin; Scannapieco, Evan; Scalo, Jon (1 October 2013). "Modeling the Pollution of Pristine Gas in the Early Universe". The Astrophysical Journal. 775 (2): 111. arXiv: 1306.4663 . Bibcode:2013ApJ...775..111P. doi:10.1088/0004-637X/775/2/111. S2CID   119233184.
  21. G. Kirchhoff (July 1860). "On the relation between the radiating and absorbing powers of different bodies for light and heat". The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science. Taylor & Francis. 20 (130).
  22. Nahar, Anil K. Pradhan, Sultana N. (2010). Atomic astrophysics and spectroscopy. Cambridge: Cambridge University Press. pp. 7, 221. ISBN   978-0-521-82536-8.{{cite book}}: CS1 maint: multiple names: authors list (link)
  23. Mahmoud Massoud (2005).  2.1 Blackbody radiation". Engineering thermofluids: thermodynamics, fluid mechanics, and heat transfer. Springer. p. 568. ISBN   3-540-22292-8.
  24. "2018 CODATA Value: Wien wavelength displacement law constant". The NIST Reference on Constants, Units, and Uncertainty. NIST. 20 May 2019. Retrieved 2019-05-20.
  25. "Luminosity of Stars". Australia Telescope National Facility. 12 July 2004. Archived from the original on 9 August 2014. Retrieved 2 July 2012.
  26. "2018 CODATA Value: Stefan–Boltzmann constant". The NIST Reference on Constants, Units, and Uncertainty. NIST. 20 May 2019. Retrieved 2019-05-20.
  27. Zwicky, F. (October 1937). "On the Masses of Nebulae and of Clusters of Nebulae". The Astrophysical Journal. 86: 217. Bibcode:1937ApJ....86..217Z. doi: 10.1086/143864 .
  28. Romanowsky, Aaron J.; Douglas, Nigel G.; Arnaboldi, Magda; Kuijken, Konrad; Merrifield, Michael R.; Napolitano, Nicola R.; Capaccioli, Massimo; Freeman, Kenneth C. (19 September 2003). "A Dearth of Dark Matter in Ordinary Elliptical Galaxies". Science. 301 (5640): 1696–1698. arXiv: astro-ph/0308518 . Bibcode:2003Sci...301.1696R. doi:10.1126/science.1087441. PMID   12947033. S2CID   120137872.
  29. Matthews, Thomas A.; Sandage, Allan R. (July 1963). "Optical Identification of 3c 48, 3c 196, and 3c 286 with Stellar Objects". The Astrophysical Journal. 138: 30. Bibcode:1963ApJ...138...30M. doi: 10.1086/147615 .
  30. 1 2 Wallace, P.R. (1991). Physics : imagination and reality. Singapore: World Scientific. pp. 235–246. ISBN   997150930X.
  31. Chiu, Hong-Yee (1964). "Gravitational collapse". Physics Today. 17 (5): 21–34. Bibcode:1964PhT....17e..21C. doi: 10.1063/1.3051610 .
  32. Rubin, Vera C.; Graham, J. A.; Kenney, Jeffrey D. P. (July 1992). "Cospatial counterrotating stellar disks in the Virgo E7/S0 galaxy NGC 4550". The Astrophysical Journal. 394: L9. Bibcode:1992ApJ...394L...9R. doi: 10.1086/186460 .
  33. Kudritzki, R.-P. (May 2010). "Dissecting galaxies with quantitative spectroscopy of the brightest stars in the Universe". Astronomische Nachrichten. 331 (5): 459–473. arXiv: 1002.5039 . Bibcode:2010AN....331..459K. doi:10.1002/asna.200911342. S2CID   119211740.
  34. 1 2 3 4 Kitchin, C.R. (1987). Stars, nebulae, and the interstellar medium : observational physics and astrophysics. Bristol: A. Hilger. pp. 265–277. ISBN   0-85274-580-X.
  35. Huggins, Sir William (1899). The Scientific Papers of Sir William Huggins. London: William Wesley and Son. pp. 114–115.
  36. 1 2 3 Tennyson, Jonathan (2005). Astronomical spectroscopy : an introduction to the atomic and molecular physics of astronomical spectra ([Online-Ausg.]. ed.). London: Imperial College Press. pp. 46–47, 99–100. ISBN   1-86094-513-9.
  37. Hirsh, Richard F (June 1979). "The Riddle of the Gaseous Nebulae". Isis. 70 (2): 162–212. Bibcode:1979Isis...70..197H. doi:10.1086/352195. JSTOR   230787. S2CID   123234614.
  38. Bowen, I. S. (1 October 1927). "The Origin of the Nebulium Spectrum". Nature. 120 (3022): 473. Bibcode:1927Natur.120..473B. doi: 10.1038/120473a0 . S2CID   4066813.
  39. Efremov, Yu. N. (22 February 2011). "On the spiral structure of the Milky Way Galaxy". Astronomy Reports. 55 (2): 108–122. arXiv: 1011.4576 . Bibcode:2011ARep...55..108E. doi:10.1134/S1063772911020016. S2CID   55372968.
  40. Shu, Frank H. (1982). The physical universe : an introduction to astronomy (12. [Dr.]. ed.). Sausalito, Calif.: Univ. Science Books. pp.  232–234. ISBN   0-935702-05-9.
  41. Hudson, Reggie L. "The Interstellar Medium". Goddard Space Flight Center Astrochemistry Laboratory. Archived from the original on 13 July 2013. Retrieved 19 November 2013.
  42. 1 2 Cami, J.; Bernard-Salas, J.; Peeters, E.; Malek, S. E. (22 July 2010). "Detection of C60 and C70 in a Young Planetary Nebula". Science. 329 (5996): 1180–1182. Bibcode:2010Sci...329.1180C. doi:10.1126/science.1192035. PMID   20651118. S2CID   33588270.
  43. Johansson, LE; Andersson, C; Ellder, J; Friberg, P; Hjalmarson, A; Hoglund, B; Irvine, WM; Olofsson, H; Rydbeck, G (1984). "Spectral scan of Orion A and IRC+10216 from 72 to 91 GHz". Astronomy and Astrophysics. 130: 227–56. Bibcode:1984A&A...130..227J. PMID   11541988.
  44. Millar, TJ; DA Williams (1993). Dust and chemistry in astronomy. Bristol [u.a.]: Inst. of Physics. p. 116. ISBN   0-7503-0271-2.
  45. "Hubble Pinpoints Furthest Protocluster of Galaxies Ever Seen". ESA/Hubble Press Release. Retrieved 13 January 2012.
  46. Haynes, Martha. "Hubble's Law". Cornell University. Retrieved 26 November 2013.
  47. Huchra, John. "Extragalactic Redshifts". California Institute of Technology. Retrieved 26 November 2013.
  48. Ellis, Richard S.; McLure, Ross J.; Dunlop, James S.; Robertson, Brant E.; Ono, Yoshiaki; Schenker, Matthew A.; Koekemoer, Anton; Bowler, Rebecca A. A.; Ouchi, Masami; Rogers, Alexander B.; Curtis-Lake, Emma; Schneider, Evan; Charlot, Stephane; Stark, Daniel P.; Furlanetto, Steven R.; Cirasuolo, Michele (20 January 2013). "The Abundance of Star-Forming Galaxies in the Redshift Range 8.5-12: New Results from the 2012 Hubble Ultra Deep Field Campaign". The Astrophysical Journal. 763 (1): L7. arXiv: 1211.6804 . Bibcode:2013ApJ...763L...7E. doi:10.1088/2041-8205/763/1/L7. S2CID   17883532.
  49. "Hubble census finds galaxies at redshifts 9 to 12". NASA/ESA. Retrieved 26 November 2013.
  50. "Planck reveals an almost perfect universe". ESA. 21 March 2013. Retrieved 26 November 2013.
  51. "Peculiar Velocity". Swinburne University of Technology. Retrieved 26 November 2013.
  52. Yasuda, Naoki; Fukugita, Masataka; Okamura, Sadanori (February 1997). "Study of the Virgo Cluster Using the B‐Band Tully‐Fisher Relation". The Astrophysical Journal Supplement Series. 108 (2): 417–448. Bibcode:1997ApJS..108..417Y. doi: 10.1086/312960 .
  53. 1 2 "Types of Binary Stars". Australia Telescope Outreach and Education. Australia Telescope National Facility. Archived from the original on 8 December 2013. Retrieved 26 November 2013.
  54. 1 2 Gray, Richard O.; Christopher J. Corbally (2009). Stellar spectral classification. Princeton, N.J.: Princeton University Press. pp. 507–513. ISBN   978-0-691-12510-7.
  55. Goody, Richard M.; Yung, Yuk Ling (1989). Atmospheric Radiation: Theoretical Basis. New York, New York, USA: Oxford University Press. ISBN   0-19-505134-3.
  56. Tessenyi, M.; Tinetti, G.; Savini, G.; Pascale, E. (November 2013). "Molecular detectability in exoplanetary emission spectra". Icarus. 226 (2): 1654–1672. arXiv: 1308.4986 . Bibcode:2013Icar..226.1654T. doi:10.1016/j.icarus.2013.08.022.
  57. Bus, S (July 2002). "Phase II of the Small Main-Belt Asteroid Spectroscopic Survey A Feature-Based Taxonomy". Icarus. 158 (1): 146–177. Bibcode:2002Icar..158..146B. doi:10.1006/icar.2002.6856. S2CID   4880578.
  58. Chapman, Clark R.; Morrison, David; Zellner, Ben (May 1975). "Surface properties of asteroids: A synthesis of polarimetry, radiometry, and spectrophotometry". Icarus. 25 (1): 104–130. Bibcode:1975Icar...25..104C. doi:10.1016/0019-1035(75)90191-8.
  59. Sekanina, Zdenek; Kracht, Rainer (3 Jun 2015). "Disintegration of Comet C/2012 S1 (ISON) Shortly Before Perihelion: Evidence From Independent Data Sets". arXiv: 1404.5968v6 [astro-ph.EP].
  60. Knight, Matthew. "Why does ISON look green?". Comet ISON Observing Campaign. Archived from the original on 3 December 2013. Retrieved 26 November 2013.
  61. Lisse, C. M.; Dennerl, K.; Englhauser, J.; Harden, M.; Marshall, F. E.; Mumma, M. J.; Petre, R.; Pye, J. P.; Ricketts, M. J.; Schmitt, J.; Trumper, J.; West, R. G. (11 October 1996). "Discovery of X-ray and Extreme Ultraviolet Emission from Comet C/Hyakutake 1996 B2". Science. 274 (5285): 205–209. Bibcode:1996Sci...274..205L. doi:10.1126/science.274.5285.205. S2CID   122700701.