nebula

nebula
nebular, adj.
/neb"yeuh leuh/, n., pl. nebulae /-lee', -luy'/, nebulas.
1. Astron.
a. Also called diffuse nebula. a cloud of interstellar gas and dust. Cf. dark nebula, emission nebula, reflection nebula.
b. (formerly) any celestial object that appears nebulous, hazy, or fuzzy, and extended in a telescope view.
2. Pathol.
a. a faint opacity in the cornea.
b. cloudiness in the urine.
3. any liquid medication prepared for use as a spray.
[1655-65; < L: a mist, vapor, cloud; akin to Gk nephéle cloud, G Nebel fog, haze]

* * *

I
Any of various tenuous clouds of gas and dust in interstellar space.

Nebulae constitute only a small percentage of a galaxy's mass. Dark nebulae (e.g., the Coalsack) are very dense, cold molecular clouds that appear as large, obscure, irregularly shaped areas in the sky. Bright nebulae (e.g., the Crab Nebula, planetary nebula) appear as faintly luminous, glowing surfaces; they emit their own light or reflect that of stars near them. The term nebula also formerly referred to galaxies outside the Milky Way Galaxy.
II
(as used in expressions)

* * *

Introduction
Latin“mist,” or “cloud”plural  nebulae,  or  nebulas 
 any of the various tenous clouds of gas and dust that occur in interstellar space. The term was formerly applied to any object outside the solar system that had a diffuse appearance rather than a pointlike image, as in the case of a star. This definition, adopted at a time when very distant objects could not be resolved into great detail, unfortunately includes two unrelated classes of objects: the extragalactic nebulae, now called galaxies, which are enormous collections of stars and gas; and the galactic nebulae, which are composed of the interstellar medium (the gas between the stars, with its accompanying small solid particles) within a single galaxy. Today the term nebula generally refers exclusively to the interstellar medium.

      In a spiral galaxy the interstellar medium makes up 3 to 5 percent of the galaxy's mass, but within a spiral arm its mass fraction increases to about 20 percent. About 1 percent of the mass of the interstellar medium is in the form of “dust”—small, solid particles that are efficient in absorbing and scattering radiation. Much of the rest of the mass within a galaxy is concentrated in visible stars, but there are strong indications from the rotation of galaxies about their centres that there is also some form of dark matter accounting for a substantial fraction of the mass in the outer regions. This material might consist of dim stars or compact objects such as neutron stars or black holes.

      The most conspicuous property of interstellar gas is its clumpy distribution on all size scales observed, from the size of the entire Milky Way Galaxy (about 1020 metres, or hundreds of thousands of light-years) down to the distance from the Earth to the Sun (less than 1011 metres, or a few light-minutes). The large-scale variations are seen by direct observation; the smallest are observed by fluctuations in the intensity of radio waves, similar to the “twinkling” of starlight caused by unsteadiness in the Earth's atmosphere. Various regions exhibit an enormous range of densities and temperatures. Within the Galaxy's spiral arms about half of the mass of the interstellar medium is concentrated in molecular clouds, in which hydrogen occurs in molecular form (H2) and temperatures are as low as 10 kelvins. These clouds are inconspicuous optically and are detected principally by their carbon monoxide (carbon dioxide) (CO) emissions in the millimetre wavelength range. Their densities in the regions studied by carbon monoxide emissions are typically 1,000 H2 molecules per cubic centimetre. At the other extreme is the gas between the clouds with a temperature of 10 million kelvins and a density of only 0.001 H+ ion per cubic centimetre. Such gas is produced by supernovas, the violent explosions of unstable stars.

      This article surveys the basic varieties of galactic nebulae distinguished by astronomers and their chemical composition and physical properties.

Classes of nebulae
      All nebulae observed in the Galaxy are forms of interstellar matter—namely, the gas between the stars that is almost always accompanied by solid grains of cosmic dust. Their appearance differs widely, depending not only on the temperature and density of the material observed but also on how the material is spatially situated with respect to the observer. Their chemical composition, however, is fairly uniform; it corresponds to the composition of the universe in general in that approximately 90 percent of the constituent atoms are hydrogen and nearly all of the rest are helium, with oxygen, carbon, neon, nitrogen, and the other elements together making up about two atoms per thousand. Based on appearance, nebulae can be divided into two broad classes: dark nebulae (dark nebula) and bright nebulae. Dark nebulae appear as irregularly shaped black patches in the sky and blot out the light of the stars lying beyond them. Bright nebulae appear as faintly luminous, glowing surfaces; they either emit their own light or reflect the light of nearby stars.

      Dark nebulae are very dense and cold molecular clouds; they contain about half of all interstellar material. Typical densities range from hundreds to millions (or more) of hydrogen molecules per cubic centimetre. These clouds are the sites where new stars are formed through the gravitational collapse of some of their parts. Most of the remaining gas is in the diffuse interstellar medium, relatively inconspicuous because of its very low density (about 0.1 hydrogen atom per cubic centimetre) but detectable by its radio emission of the 21-centimetre line of neutral hydrogen.

      Bright nebulae are comparatively dense clouds of gas within the diffuse interstellar medium. They have several subclasses: (1) reflection nebulae, (2) diffuse nebulae, (3) planetary nebulae, (4) supernova remnants, and (5) diffuse ionized gas.

      Reflection nebulae reflect the light of a nearby star from their constituent dust grains. The gas of reflection nebulae is cold, and such objects would be seen as dark nebulae if it were not for the nearby light source.

      Diffuse nebulae are clouds of hydrogen ionized (separated into positive H+ ions and free electrons) by a neighbouring hot star. The star must be of stellar type O or B, the most massive and hottest of normal stars in the Galaxy, in order to produce enough of the radiation required to ionize the hydrogen.

 Planetary nebulae (planetary nebula) are composed of the outer parts of stars that are dying but are not massive enough to become supernovas—namely, red giant stars. That is to say, a red giant has shed its outer envelope in a less violent event than a supernova explosion and has become an intensely hot star surrounded by a shell of material that is expanding at a speed of tens of kilometres per second. Planetary nebulae typically appear as rather round objects of relatively high surface brightness. Their name is derived from their superficial resemblance to planets—i.e., their regular appearance when viewed telescopically as compared to the chaotic forms of other types of nebula.

      Supernova remnants are the clouds of gas expanding at speeds of hundreds or even thousands of kilometres per second from comparatively recent explosions of massive stars. If a supernova remnant is younger than a few thousand years, it may be assumed that the gas in the nebula was mostly ejected by the exploded star. Otherwise, the nebula would consist chiefly of interstellar gas that has been swept up by the expanding remnant of older objects.

      Diffuse ionized gas, so pervasive among the nebular clouds, has been recently recognized as a major component of the Galaxy. It is observed by faint emissions of H+, N+, and S+ (positive hydrogen, nitrogen, and sulfur ions) detectable in all directions, collectively requiring far more power than the much more spectacular diffuse nebulae, planetary nebulae, or supernova remnants that occupy a tiny fraction of the volume. The origin of the diffuse ionized gas is still somewhat controversial (see below Diffuse ionized gas (nebula)).

Historical survey of the study of nebulae

Pre-20th-century observations of nebulae
 In 1610, two years after the invention of the telescope, the Orion Nebula, which looks like a star to the naked eye, was discovered by the French scholar and naturalist Nicolas-Claude Fabri de Peiresc (Peiresc, Nicolas-Claude Fabri de). In 1656 Christiaan Huygens (Huygens, Christiaan), the Dutch scholar and scientist, using his own greatly superior instruments, was the first to describe the bright inner region of the nebula and to determine that its inner star is not single but a compact quadruple system.

      Early 18th-century observational astronomers gave high priority to comet seeking. A by-product of their search was the discovery of many bright nebulae. Several catalogs of special objects were compiled by comet researchers; by far the best known is that of the Frenchman Charles Messier (Messier catalog) who in 1781 compiled a catalog of 103 nebulous, or extended, objects in order to prevent their confusion with comets. Most are clusters of stars, 35 are galaxies, and 11 are nebulae. Even today many of these objects are commonly referred to by their Messier catalog number: M20, for instance, is the great Trifid Nebula, in the constellation Sagittarius.

The work of the Herschels
      By far the greatest observers of the early and middle 19th century were the English astronomers William Herschel (Herschel, Sir William) and his son John (Herschel, Sir John, 1st Baronet). Between 1786 and 1802 William Herschel, aided by his sister Caroline (Herschel, Caroline Lucretia), compiled three catalogs totaling about 2,500 clusters, nebulae, and galaxies. John Herschel later added to the catalogs 1,700 other nebulous objects in the southern sky visible from the Cape Observatory in South Africa but not from London and 500 more objects in the northern sky visible from England.

      The catalogs of the Herschels formed the basis for the great New General Catalogue (New General Catalogue of Nebulae and Clusters of Stars) (NGC) of J.L. Dreyer, published in 1888. It contains the location and a brief description of 7,840 nebulae, galaxies, and clusters. In 1895 and 1908 it was supplemented by two Index Catalogues (IC) of 5,386 additional objects. The list still included galaxies as well as true nebulae, for they were often at this time still indistinguishable. Most of the brighter galaxies are still identified by their NGC or IC numbers according to their listing in the New General Catalogue or Index Catalogues.

Advances brought by photography and spectroscopy
      The advent of photography, which allows the recording of faint details invisible to the naked eye and provides a permanent record of the observation for study of fine details at leisure, caused a revolution in the understanding of nebulae. In 1880 the first photograph of the Orion Nebula was made, but really good ones were not obtained until 1883. These early photographs showed a wealth of detail extending out to distances unsuspected by visual observers.

      Much can be learned about the physical nature of an astronomical object by studying its spectrum (spectroscopy)—i.e., the resolution of its light into different wavelengths (or colours). Study of the spectrum of an object provides a decisive test as to whether it is composed of unresolved stars (as are galaxies) or glowing gas. Stars radiate at almost all wavelengths, with perhaps dark absorption lines superimposed, while hot, transparent gas clouds radiate only emission lines at certain wavelengths characteristic of their constituent gases. In 1864 observation of the spectrum of the Orion Nebula showed bright emission lines of glowing gases, with conspicuous hydrogen lines and some green lines even brighter. By contrast, the spectrum of galaxies was found to be stellar, so that a distinction between galaxies and nebulae—that nebulae are gaseous and galaxies are stellar—was appreciated at this time, although the true sizes and distances of galaxies were not demonstrated until the 20th century.

20th-century discoveries
      The 20th century has witnessed enormous advances in observational techniques as well as scientific understanding of the physical processes that operate in interstellar matter.

      In 1930 a German optical worker, Bernhard Schmidt (Schmidt, Bernhard Voldemar), invented an extremely fast wide-angled camera ideal for photographing faint, extended nebulae. In addition, photographic plates became progressively more sensitive to an ever-widening range of colours. Since about 1960, however, photography has been almost completely replaced for research purposes by photoelectric devices. Most images are now recorded with so-called charge-coupled devices (or CCDs) that act as arrays of tiny photoelectric cells, each recording the light from a small patch of sky. Modern CCDs consist of square arrays of up to 4,000 cells on each side, or 16 million independent photocells, capable of observing the sky simultaneously. CCDs possess three advantages over photography: (1) their sensitivity to light is up to 100 times greater; (2) the range of light levels that they can record in one exposure is much greater, so that information about both the bright and faint parts of the same nebula can be obtained in a single exposure; and (3) they (or similar arrays) can record light ranging from 0.1 micrometre in the vacuum ultraviolet (accessible only from satellites orbiting above the Earth's atmosphere) to more than 1.2 micrometres in the infrared.

      The most important recent development for the study of nebulae has been the use of spacecraft, which allows the observation of radiation normally absorbed by the Earth's atmosphere: X rays (which have very short wavelengths), far-ultraviolet radiation (with wavelengths shorter than about 0.3 micrometre, below which atmospheric ozone is strongly absorbing), and infrared (from about 3 micrometres to 1 mm) strongly absorbed by atmospheric water vapour and carbon dioxide. X rays and ultraviolet radiation (ultraviolet astronomy) reveal the physical conditions in the hottest regions in space (extending to some 100 million kelvins in shocked supernova gas). Infrared radiation (infrared astronomy) reveals the conditions within dark, cold clouds, into which starlight cannot penetrate because of absorbing dust layers.

      The primary means of studying nebulae is not by images but by spectra, by means of prisms (in the earlier part of the century), diffraction gratings, or crystals in the case of X rays. A particularly useful instrument is the echelle spectrograph, in which one coarsely ruled grating spreads the electromagnetic radiation in one direction, while another finely ruled grating disperses it in the perpendicular direction. This device, often used both in spacecraft and on the ground, allows astronomers to record simultaneously a wide range of wavelengths with very high spectral resolution (i.e., to distinguish slightly differing wavelengths). For even higher spectral resolution astronomers employ Fabry-Pérot interferometers. Spectra provide powerful diagnostics of the physical conditions within nebulae. Images and spectra provided by Earth-orbiting satellites, especially the Hubble Space Telescope, have yielded data of unprecedented quality.

      Ground-based observations also have played a major role in recent advances in scientific understanding of nebulae. The emission of gas in the radio and submillimetre ranges of wavelengths provides crucial information regarding the physical conditions and molecular composition of gas. Large radio telescope arrays (radio telescope), in which several individual telescopes function collectively as a single enormous instrument, give spatial resolutions in the radio regime vastly better than any yet achieved by optical means.

Chemical composition and physical properties of nebulae
      Many characteristics of nebulae are determined by the physical state of their constituent hydrogen, by far the most abundant element. For historical reasons, nebulae in which hydrogen is mainly ionized (H+) are called H II regions (H II region); those in which hydrogen is mainly neutral are designated H I regions (hydrogen cloud); and those in which the gas is in molecular form (H2) are referred to as molecular clouds. The distinction is important because neutral hydrogen atoms are extremely efficient at absorbing ionizing radiation, with an energy per photon of at least 13.6 electron volts or, equivalently, a wavelength of 0.0912 micrometre or less. If the hydrogen is mainly neutral, no ionizing radiation can penetrate except for photons with energies in the X-ray range and above, or thousands of electron volts or more, in which case the hydrogen becomes somewhat transparent. For photon energies less than 13.6 volts, the radiation within an H I region has a spectrum similar to that of a fairly hot star (i.e., about 15,000 K), since hot stars produce much more radiation than do cooler ones. However, the absorption by neutral hydrogen abruptly reduces the radiation field to almost zero for energies above 13.6 volts. This dearth of hydrogen-ionizing radiation within the H I region implies that no ions requiring more energy than hydrogen can be produced, and the ionic species of all elements are limited to the lower stages of ionization. Within H II regions, with almost all of the hydrogen ionized and thereby rendered nonabsorbing, photons of all energies propagate, and ions requiring energetic radiation for their production (e.g., O++) occur.

      Ultraviolet photons (ultraviolet radiation) of more than 11.2 volts can dissociate molecular hydrogen (H2) into two H atoms. In H I regions there are enough of these photons to prevent the amount of H2 from becoming large, but the destruction of H2 as fast as it forms takes its toll on the number of photons of suitable energies. Furthermore, interstellar dust is a fairly efficient absorber of photons throughout the optical and ultraviolet range. In some regions of space the number of photons with energies higher than 11.2 volts is reduced to the level where H2 cannot be destroyed as fast as it is produced on grain surfaces. In this case, H2 becomes the dominant form of hydrogen present. The gas is then part of a molecular cloud. The role of interstellar dust in this process is crucial because H2 cannot be formed efficiently in the gas phase.

Interstellar dust
      Only about 0.7 percent of the mass of the interstellar medium is in the form of solid grains (interplanetary dust particle), but these grains have a profound effect on the physical conditions within the gas. Their main effect is to absorb stellar radiation; for photons unable to ionize hydrogen and for wavelengths outside of absorption lines or bands, the dust grains are much more opaque than the gas. The dust absorption increases with photon energy, so that long-wavelength radiation (radio and far-infrared) can penetrate dust freely, near-infrared rather well, and ultraviolet relatively poorly. Dark, cold clouds, within which all star formation takes place, owe their existence to dust. Besides absorbing starlight, the dust acts to heat the gas under some conditions (by ejecting electrons produced by the photoelectric effect, following the absorption of a stellar photon) and to cool the gas under other conditions (because the dust can radiate energy more efficiently than the gas, and so in general is colder). The optical appearance of most nebulae is significantly modified by the obscuring effects of the dust.

      The chemical composition of the gas phase of the interstellar medium alone, without regard to the solid dust, can be determined from the strength of narrow absorption lines in the spectra of background stars that are produced by the gas. Comparison of the composition of the gas with cosmic (solar) abundances shows that almost all of the iron, magnesium, and silicon, much of the carbon, and only some of the oxygen and nitrogen are contained in the dust. The absorption and scattering properties of the dust reveal that the solid grains are composed partially of silicaceous material similar to terrestrial rocks, though of an amorphous (amorphous solid) rather than crystalline variety; the grains also have a carbonaceous component. The carbon dust probably occurs in at least two forms: (1) graphitic grains, produced by the temporary heating of very small amorphous carbon particles after the absorption of a single photon of ultraviolet light, and (2) individual freely floating hydrocarbon molecules, with a range varying from 70 to several hundred carbon atoms and some hydrogen atoms that dangle from the outer edges of the molecule or are trapped in the middle of it. It is merely convention that these molecules are referred to as dust, since the smallest may be only somewhat larger than the largest molecules observed with a radio telescope. Both of the dust components are needed to explain spectroscopic features arising from the dust. In addition, there are probably mantles of hydrocarbon on the surfaces of the grains. The size of the grains ranges from perhaps as small as 0.0003 micrometre for the tiniest hydrocarbon molecules to a substantial fraction of a micrometre; there are many more small grains than large ones.

      The dust cannot be formed directly from purely gaseous material at the low densities found even in comparatively dense interstellar clouds, which would be considered an excellent laboratory vacuum. For a solid to condense, the gas density must be high enough to allow a few atoms to collide and stick together long enough to radiate away their energy to cool and form a solid. Grains are known to form in the outer atmospheres of cool supergiant (supergiant star) stars where the gas density is comparatively high (perhaps 109 times what it is in such nebulae). The grains are then blown out of the atmosphere by radiation pressure (the mechanical force of the light they absorb and scatter). Calculations indicate that refracting materials, such as the constituents of the grains proposed above, should condense in this way.

      There is clear indication that the dust is heavily modified within the interstellar medium by interactions with itself and with the interstellar gas. The absorption and scattering properties of dust show that there are many more smaller grains in the diffuse interstellar medium than in dense clouds; apparently in the dense medium the small grains have coagulated into larger ones, thereby lowering the ability of the dust to absorb radiation with short wavelengths (namely, ultraviolet, near 0.1 micrometre). The gas-phase abundances of some elements, such as iron, magnesium, and nickel, also are much lower in the dense regions than in the diffuse gas, although even in the diffuse gas most of these elements are missing from the gas and are therefore condensed into dust. These systematic interactions of gas and dust show that dust grains collide with gas atoms much more rapidly than one would expect if the dust and gas simply drifted together. There must be disturbances, probably magnetic in nature, that keep the dust and gas moving with respect to each other.

Dark nebulae (dark nebula)
      The dark nebulae are clumps or clouds that become opaque because of their internal dust grains. The form of such dark clouds is very irregular: they have no clearly defined outer boundaries and sometimes take on convoluted serpentine shapes. The largest dark nebulae are visible to the naked eye, appearing as dark patches against the brighter background of the Milky Way. An example is the Coalsack in the southern sky.

      The hydrogen of these opaque dark clouds exists in the form of H2 molecules. The largest nebulae of this type, the so-called giant molecular clouds, are more than a million times as massive as the Sun. They contain much of the mass of the interstellar medium, are some 150 light-years across, and have an average density of 100 to 300 molecules per cubic centimetre and an internal temperature of only 7 to 15 K. Molecular clouds consist mainly of gas and dust but contain many stars as well. The cloud cores are completely hidden from view and would be undetectable except for the microwave emissions from their constituent molecules. This radiation is not absorbed by dust and readily escapes the cloud. The material within the clouds is clumped together in all sizes, with some clouds ranging down to the masses of individual stars. The density within the clumps may reach up to 105 H2 per cubic centimetre or more. Small clumps may extend about one light-year across. The clouds have an internal magnetic field that provides support against their own gravity.

      As one penetrates into the interior of a molecular cloud, the chemistry and physical conditions become quite different from those of the surrounding low-density interstellar medium. In the outer parts of the dark cloud, the hydrogen is neutral. Deeper within it, as dust blocks out an increasing amount of stellar ultraviolet radiation, the cloud becomes darker and colder. As the centre is approached, the predominant form of carbon changes successively from C+ (on the outside) to neutral C to finally the molecule carbon monoxide (CO), which is so stable that it remains the major form of carbon in the gas phase in the darkest regions. At great depths within the cloud other molecules can be seen from their microwave transitions, and more than 70 chemical species have been identified within the constituent gas. Owing to the comparatively low densities and low temperatures, the chemistry is very exotic, as judged by terrestrial experiments; some rather unstable species can exist in space because there is not enough energy to convert them to more stable forms. An example is the near equality of the abundances of the interstellar molecule HNC (hydroisocyanic acid) and its isomer HCN (hydrocyanic acid); in ordinary terrestrial conditions there is plenty of energy to allow the nitrogen and carbon atoms in HNC to exchange positions and produce HCN, by far the preferred species for equilibrium chemistry. In the cold clouds, however, not enough energy exists for the exchange to occur. There is less than one-thousandth as much starlight within a cloud as in the interstellar space outside the cloud, and the heating of the material in the cloud is provided primarily by cosmic rays (cosmic ray). Cooling within the cloud occurs chiefly by transitions between low-lying levels of the carbon monoxide molecule.

      Recent observations of the emission lines from C+, C0, and CO show that the edges of the dark nebulae are very convoluted spatially, with stellar ultraviolet radiation able to penetrate surprisingly far throughout the cloud despite the absorption of the dust. Stellar radiation can apparently enter the cloud through channels where the dust (and gas) density is lower than average. The clumpiness of the interstellar material has profound effects on its properties.

      In the inner regions of dark nebulae important events take place: the formation of stars. The discussion that follows will be slightly oversimplified by the assumption that a nebula has very little net rotation. Since the density in a star is immensely greater than that in a nebula, star formation must occur through condensation. The force of gravity (gravitation) in the nebula constantly pulls it together, working against the disruptive collisions of any one nebula or cloud with its neighbours and against the pressure provided by an internal magnetic field. But, even if gravity can hold the cloud together against other forces, the cloud cannot collapse unless it can cool; gravitational energy is released in a contraction and would heat the gas cloud, thus increasing the outward pressure and preventing further collapse. The dust grains are efficient emitters of infrared radiation, which escapes, removing the energy from the nebula. A cloud, therefore, contracts under its own gravity, radiating away half of the gravitational energy of the contraction, the other half going into heating the gas. As it contracts, the density and gravitational binding increase until finally the cloud's gravity so dominates the internal pressure that the material rushes inward in almost a free fall.

      While the entire cloud has been collapsing, it does not have a smooth density distribution but rather consists of a chaotic jumble of smaller clouds. These cloudlets pull themselves together by their own gravity into “protostars,” each of which is destined to be an individual stellar system. Most of these protostars are smaller than one solar mass, but a very few may be several (up to about 100) times as massive as the Sun. These few massive stars have a profound influence in the evolution of the nebula.

      Each protostar collapses very quickly; its gas falls inward in free fall. A protostar can collapse from a size equal to the outer diameter of the solar system to about 30 times the Sun's size (the size of Mercury's orbit) in about six months. After it reaches that size, the material becomes so hot it has become a star—that is, an opaque body radiating energy only from its surface. Around it is a whirling ring of cold, dusty material, which will also break up into even smaller fragments, the “protoplanets.” (protoplanet) It is now believed that planetary systems are made out of this dusty chaos.

      Such newly formed stars continue to contract, becoming hotter but less luminous in the process, until they start to produce heat by converting hydrogen into helium rather than by contracting. If massive enough, a star ionizes the nebular material, producing a bright nebula around it.

      These ideas are given encouraging confirmation by observations of dark nebulae in very long wavelength infrared radiation. Some of the brightest infrared sources are associated with such dark dust clouds; a good example is the class of T Tauri variables (T Tauri star), named for their prototype star in the constellation Taurus. The T Tauri stars are known for a variety of reasons to be extremely young. The variables are always found in or near dark nebulae; they often are also powerful sources of infrared radiation, corresponding to warm clouds of dust heated by the T Tauri star to a few hundred kelvins. There are some strong infrared sources (especially in the constellation of Orion) that have no visible stars with them; these are presumably “cocoon stars” completely hidden by their veils of dust.

Neutral clouds (hydrogen cloud) (H I regions)
      Neutral hydrogen is dominant in clouds that have enough starlight to dissociate molecular hydrogen but lack hydrogen-ionizing photons from hot stars. These clouds can be seen as separate structures within the lower-density interstellar medium or else on the outer edges of the molecular clouds. Because a neutral cloud moves through space as a single entity, it often can be distinguished by the absorption line that its atoms or ions produce at their common radial velocity in the spectrum of a background star.

      Neutral clouds have two possible temperatures (with little material at intermediate values), both determined by the balance of heating and cooling rates. Cold H I regions are at about 80 K; heating is provided by electrons ejected from the dust grains by interstellar ultraviolet radiation incident upon such a cloud from outside. Cooling is mainly by + (carbon) because passing electrons or hydrogen atoms can excite it from its normal energy state, the lowest, to one slightly higher, followed by emission of radiation at 158 micrometres. This line is observed to be very strong in the spectrum of the Galaxy as a whole, indicating that a great deal of energy is removed from interstellar gas by this process. Warm H I regions (about 8,000 K) are cooled by excitation of the n = 2 level of hydrogen, which is at a much higher energy than the lowest level of C+ and therefore requires a higher temperature for its excitation. The density is much lower in the hotter regions, rendering the cooling by C+ less efficient because it requires collisions for the excitation. At any particular density there is far more neutral hydrogen available for cooling than C+. Clouds of either temperature can exist at the same pressure, so the interstellar medium separates itself into the two phases depending on the past history of each parcel of gas.

Reflection nebulae
 If a cloud that would normally be a dark nebula happens to be illuminated by a nearby bright star not hot enough to ionize the cloud's hydrogen, the dust grains in the cloud reflect the starlight and give rise to a reflection nebula. The famous nebulosity in the Pleiades star cluster is of this type; it was discovered in 1912 that the spectrum of this nebula mimics the absorption lines of the nearby stars, whereas bright nebulae that emit their own light show characteristic emission lines quite unlike stars. The brightest reflection nebulae are illuminated by B-type stars that are very luminous but have temperatures lower than about 25,000 K, somewhat cooler than the O-type stars that would ionize the hydrogen in the gas and produce a diffuse emission nebula. The extent and brightness of reflection nebulae show conclusively that dust grains are excellent reflectors in the broad range of wavelengths extending from the ultraviolet (as determined from observations from space) through the visible. Optical observations suggest that about 60–70 percent of the light is reflected rather than absorbed, while the corresponding fraction for the whole Earth is only 35 percent and for the Moon a mere 5 percent. Grains reflect light almost as well as fresh snow, more because of their favourable size (which promotes scattering rather than absorption) than their chemical composition. Calculations show that even graphite, which is black in bulk, reflects visible light well when dispersed into small particles.

Diffuse nebulae (H II regions)
  As noted above, clouds of gas and dust that contain stars hot enough to ionize their hydrogen are called diffuse nebulae, or H II regions. In contrast to dark or reflection nebulae, their spectrum consists of emission lines at various wavelengths characteristic of their ions. Like dark clouds, diffuse nebulae typically have little regular structure or sharp boundaries—hence their name. Their sizes and masses vary widely. There is even a faint region of ionized gas around the Sun and other comparatively cool stars, but it cannot be observed from nearby stars with existing instruments. The largest diffuse nebulae (none of which occur in the Milky Way Galaxy) are 500 light-years across and contain at least 100,000 solar masses of ionized gas. These enormous H II regions are powered by clusters of massive hot stars rather than by any single stellar body. A typical diffuse nebula within the Galaxy measures about 30 light-years in diameter and has an average density of about 10 atoms per cubic centimetre. The mass of such a cloud amounts to several hundred solar masses. The only diffuse nebula visible to the naked eye is the beautiful Orion Nebula (see photograph—>). Located in the constellation named for the Greek mythological hunter, it is seen as the central “star” in Orion's sword. The entire constellation is enveloped in faint emission nebulosity, powered by several stars in Orion's belt rather than by the star exciting the much smaller Orion Nebula itself. The largest diffuse nebula in terms of angular size is the Gum Nebula, named after its discoverer, the Australian astronomer Colin S. Gum. It measures 40° in angular diameter and is mainly ionized by two very hot stars (Zeta Puppis and Gamma Velorum).

      The spectra of H II regions show bright emission lines arising from two fundamentally different processes. The first are recombination lines, which are produced when any ion combines with an electron to form a neutral atom (or lower stage of ionization), radiating away some or all of the energy it had from its previous ionization. All elements produce these lines, but the lines are bright only for recombination of the very abundant ions H+ and He+. The observed lines are produced when the newly formed atom is in an excited energy level and cascades from that level to a lower one with the emission of a line photon. One of the brightest lines in diffuse nebulae, for example, is the red α line of hydrogen, arising either from recombination of an H+ ion to the n = 3 level of a hydrogen atom or from a recombination to a higher level followed by a cascade to the n = 3 level. A hydrogen atom in the n = 3 level can undergo a transition to the n = 2 level with the emission of the Hα photon.

      The second are what are termed forbidden lines, which are produced by the excitation of an atom or ion from its lowest level to a higher level by a collision with an electron, followed by the emission of a photon from the higher level downward to a lower one (not necessarily the lowest). The name forbidden is appropriate because these emission lines are not observed directly in the laboratory; the downward transition that produces the photon requires a relatively long time to take place, and in the laboratory the excited atom encounters another particle or the container walls before it emits the photon. This encounter allows the excited atom to give up its energy to its collision partner without radiating a photon. In a diffuse nebula it has time to produce the radiation before it can collide with another atom or ion. The strongest lines in most diffuse nebulae are the forbidden lines from either O++ (at 4959 and 5007 angstroms) or from O+ (at 3726 and 3729 angstroms). These lines provide very strong cooling above about 10,000 K and act to limit the temperatures of H II regions to approximately that value, unless the abundance of oxygen is significantly lower than solar abundance, in which case higher temperatures can be reached.

 Diffuse nebulae are almost always accompanied by dark nebulae (dark nebula) on their borders. The Orion Nebula, for example, is merely a conspicuous ionized region on the nearby face of a much larger dark cloud; the diffuse nebula is almost entirely produced by the ionization provided by a single hot star, one of the four bright central stars (the Trapezium) identified by Huygens in 1656. The shape of the Orion Nebula appears at visible wavelengths as irregular. However, much of this seeming chaos is spurious, caused by obscuration of dust in foreground neutral material rather than by the actual distribution of ionized material. Radio waves can penetrate the dust unhindered, and the radio emission from the ionized gas reveals it to be quite circular in shape and surprisingly symmetrical as seen in projection on the sky. The foreground dark material obscures about half of the ionized nebula.

      In most cases, a diffuse nebula is found on the outer edge of a large molecular cloud in which star formation is occurring, induced by the very presence of the diffuse nebula. For instance, behind the bright Orion Nebula, deeper within the dark, cold Orion molecular cloud, new stars are being formed today. At present, none of the new stars is massive and hot enough to produce its own diffuse nebula, but presumably one of them eventually will be. When a diffuse nebula is produced from cold molecular gas by the formation of a hot star, the temperature is raised from roughly 25 to 8,000 K, and the number of particles per cubic centimetre is almost quadrupled because each H2 molecule is split into two ions and two electrons. Gas pressure is proportional to the product of the temperature and number of particles per cubic centimetre (regardless of their mass, so electrons are as important as the much heavier ions). Thus, the pressure in a diffuse nebula is some 800 times the pressure of the cold gas from which it formed. The excess pressure causes a violent expansion of the gas into the dense cloud. Rapid star formation may occur in the compressed region, producing an expanding group of young stars. Such groups, the so-called O Associations (with O stars) or T Associations (with T Tauri stars), have been observed. The component stars simultaneously generate extremely fast outflows from their atmospheres. These winds create regions of hot, tenuous gas surrounding the association. Eventually the massive stars in the association explode as supernovas, which further disturb the surrounding gas.

      This picture of the evolution of diffuse and dark nebulae is one of constant turmoil, a few transient O stars serving to keep the material stirred, in constant motion, continually producing new stars and churning clouds of gas and dust. In this way some of the stellar thermonuclear energy is converted into the kinetic energy of interstellar gas.

Ultracompact H II region
      Recently “ultracompact” H II regions have been discovered inside the molecular clouds in which they were formed. These nebulae are observed only at the wavelengths of radio and far-infrared radiation, which are able to penetrate the thick dust in the clouds. They are extremely bright at wavelengths of 50 micrometres, and the most luminous ones were detected throughout the Galaxy by the orbiting Infrared Astronomical Satellite (IRAS), which in 1983 mapped almost the entire sky at 12, 25, 60, and 100 micrometres. There are about 2,500 of these objects in the Galaxy, representing 10 to 20 percent of the total O-type star population. Usually only a light-month in size, 100 times smaller than a typical diffuse nebula, they show densities in the ionized region of 105 hydrogen atoms per cubic centimetre. They cannot be at rest with respect to the surrounding gas; if they were, the immense pressure exerted by their dense, hot gas would cause a violent expansion. (Their lifetimes would be only about 3,000 years—exceedingly short on an astronomical time scale—and not nearly as many could be seen as the number observed by astronomers.) Rather, their gas is kept confined because they are moving through the surrounding cloud at speeds of about 10 km/s (kilometres per second), and what is observed is the cloud of freshly ionized gas ahead of them that has not yet had time to expand. The ultracompact diffuse nebulae leave behind a trail of ionized material that is not as bright as the confined gas ahead of them. This trail gradually fades as it recombines after the ionizing star has passed. The radio radiation is produced by the ionized gas, but the far-infrared radiation is emitted by the 5,000 solar masses of surrounding dust warmed by the luminosity of the embedded star.

      High-resolution studies of diffuse nebulae reveal one of the surprises that make the study of astrophysics delightful. Instead of the smooth structure that might be expected of a gas, a delicate tracery of luminous filaments can be detected, down to the smallest scale that can be resolved. In the Orion Nebula this is about 10 times the diameter of the orbit of Pluto around the Sun. Even finer detail may exist, and there is evidence from spectra that much of the matter may be gathered into dense condensations, or knots, the rest of the space being comparatively empty. The time it would take unrestrained gas to fill a vacuum between the visible filaments is only about 200 years, much less time than the age of the nebula. The nebular gas is presumably restrained from expansion by the pressure of tenuous material, which has a temperature of a few million kelvins. Its pressure, however, is comparable to that in the visible “warm” (8,000 K) gas of the diffuse nebula. Hence, the density of the hot material is several hundred times lower, which effectively prevents it from radiating. Moreover, the material is invisible except in low-energy X rays. The space throughout the plane of the Galaxy is largely filled with this hot component, mainly produced and heated from supernovas. In diffuse nebulae, it also arises from the “wind” blown off the atmospheres of the exciting stars at speeds of up to 3,000 km/s. This stellar wind creates a large cavity or bubble in the denser, cooler gas originally surrounding such a star. In the interior of the bubble, the radially flowing stellar wind passes through a transition in which its radial motion is converted into heat. The hot gas then fills most of the cavity (perhaps 90 percent or more) and serves to separate the filaments of the warm, comparatively dense diffuse nebula. Within the condensations of visible plasma, there are neutral globules in which the gas is quite cold (about 100 K) and dense enough (typically, 10,000 atoms per cubic centimetre) to have about the same pressure as the hot and warm materials. In short, a diffuse nebula is much more complicated than its visual radiation would suggest.

Supergiant nebulae
 The most energetic diffuse nebulae in the Galaxy are revealed only by radio telescope—they are so heavily obscured by dust that they are inconspicuous optically. These nebulae have over 100 times more ionizations per second than does the Orion Nebula, too many to be provided by a single star (unless there are stars more luminous than any known within the Galaxy). Indeed, very high resolution images of the most luminous nebulae show there are clumps of ionized gas—probably ionized by tight groupings of single stars—embedded in rather diffuse gas that emit most of the radio radiation. Similar objects, about 200 light-years in diameter, are formed in other galaxies; if they were located at the Orion Nebula, they would cover the entire constellation of Orion with brightly glowing gas. These supergiant nebulae are more than 10 times as luminous as any in the Galaxy. The entire Local Group—the cluster of galaxies consisting of the Galaxy, the great spiral galaxy in Andromeda, the smaller spiral in Triangulum, and more than 20 other stellar assemblages—contains but one supergiant nebula: the object called 30 Doradus, in the Large Magellanic Cloud. This nebula requires over 1,000 times more ionizations per second than the Orion Nebula. It contains a stellar cluster called R136, the source of most of the energy radiated by the nebula. This grouping consists of dozens of the most massive known stars of the Galaxy, all packed into a volume only a thousandth of a typical stellar spacing in size. How such a cluster could form, in either case, is a fascinating puzzle. There are other supergiant nebulae outside of the Local Group, some of which radiate 10 times the energy of 30 Doradus.

Location of diffuse and dark nebulae in the Galaxy
      One of the remarkable features of diffuse and dark nebulae is their concentration in the spiral arms in the plane of the Galaxy. While there is no definite boundary to the arms, which have irregularities and bifurcations, the nebulae in other spiral galaxies are strung out along these narrow lanes and form a beautifully symmetric system when viewed from another galaxy. The nebulae are remarkably close to the galactic plane; most are within 300 light-years, only 1 percent of the Sun's distance from the centre. The details of the explanation of why the gas is largely confined to the spiral arms is beyond the scope of this article (see Milky Way Galaxy: Major components (Milky Way Galaxy)). Briefly, the higher density of the stars in the arms produces sufficient gravity to hold the gas to them.

      The question may be asked: why doesn't the gas simply condense into stars and disappear? The present rate of star formation is about one solar mass per year in the entire Galaxy, which contains something like 2 × 109 solar masses of gas. Clearly, if the gas received no return of material from stars, it would be depleted in roughly 2 × 109 years, about one-fifth the present age of the Galaxy. Several processes by which gas is returned to the interstellar medium from stars have been observed. Possibly the most important is the ejection of planetary nebula shells, discussed below; other processes are ejection of material from massive O- and B-type normal stars or from cool M giants and supergiants. This rate of gas ejection is roughly equal to the rate of star formation, so that the mass of free gas is declining very slowly.

      This cycling of the gas through stars presumably has had one major effect: the chemical composition of the gas has been changed by the nuclear reactions inside the stars. There is excellent evidence that the Galaxy originally consisted of 77 percent hydrogen by mass and that almost all of the rest of the constituent matter was helium. All heavy elements have been produced inside stars by being subjected to the exceedingly high temperatures and densities in the central regions. Thus, most of the atoms and molecules on the Earth, as well as in human bodies, owe their very existence to processes that occur within stars.

Chemical composition of diffuse nebulae
      The composition of diffuse nebulae can be estimated by relating the strengths of the emission lines found in their spectra to the numbers of atoms producing them. Great strides have been made in calculating the necessary atomic properties. The principal difficulties in determining the abundances of elements from nebular emission lines are (1) the estimation of the nebular temperature, which is necessary because line emission for a given abundance increases rapidly with increasing temperature, and (2) the estimation of the abundances of stages of ionization of the element, if these stages do not have any observable lines. There are at least two reliable indicators of nebular temperatures. The first uses the relative strengths of lines arising from two or more upper levels of ions of elements heavier than helium. The levels must be reasonably different in energy. The ratio of line strengths depends sensitively on the electron temperature. Unfortunately, there are not many such diagnostic line pairs. The most frequently available pair is that of doubly ionized oxygen. The line from the upper level, at a wavelength of 4363 angstroms, is very faint, often less than 1 percent of the strong green line at 5007 angstroms. If there are fluctuations of temperature within a nebula, this method estimates the temperature to be too high. The second method of determining nebular temperature is based on the strengths of emission lines of hydrogen in very high quantum states, such as the transition between the 110 and 109 levels. Because the higher levels are closely spaced, the transitions between two higher levels have low energies and correspondingly long wavelengths—six centimetres, in this case. This radio wavelength implies that the line can be observed through large amounts of dust. There is some action originating in nebulae to amplify the emission by means of a mechanism fundamental to the laser and maser—namely, stimulated emission. The strength of a radio line depends on the nebular temperature. If there are regions of various temperatures within a nebula, such radio lines, in contrast to the forbidden lines, give an estimate that is lower than the true average. The two methods of estimating nebular temperature yield gratifyingly similar results when they are applied to the same object. This fact indicates that temperature fluctuations within a nebula may not be a serious problem.

      All abundant chemical elements have some stages of ionization that are associated with observable emission lines from which the abundance of a given ion can be determined after the temperature has been estimated in the manner discussed above. The primary interest, though, is in the total abundance of the element and not simply that of an individual stage of ionization. Ions that have no observable lines are accounted for by theoretical calculations. Elaborate computer calculations predict the ionization structure of gas ionized by a hot star; the temperature of the star is determined by matching the observed stages of ionization with the computer model. The calculations then provide predictions of the abundances of the invisible ions. The correction for unobserved stages is of little significance for some elements (oxygen) but absolutely crucial for others (argon and carbon). The final estimates for—in order of decreasing reliability—oxygen, sulfur, nitrogen, and neon abundances in diffuse nebulae are comparable in accuracy to determinations of stellar composition. Those for carbon and argon, however, are more problematic. The well-established abundances have uncertainties of about 30 percent. Such determinations apply only to the portions of the elements in the gas phase. Solids (dust grains) do not produce emission lines.

      In the Orion Nebula, the abundances of elements other than hydrogen are (in atoms per million hydrogen atoms) as follows: helium, 80,000; oxygen, 400; carbon, 320; neon, 70; nitrogen, 50; sulfur, 12; and argon, 4. One of the most enigmatic results of the Orion investigations is that the oxygen abundance (oxygen) in the nebula is only about 0.6 that in the Sun. This finding is most unexpected because supernovas presumably have been adding oxygen to the interstellar gas ever since the Sun formed some five billion years ago. Grains are probably not responsible for the “missing” oxygen in the Orion Nebula because the inner part of the nebula seems to be free of dust. Furthermore, nitrogen and neon, both unlikely to be found in dust grains, also are deficient in Orion in exactly the same proportion as oxygen. One possibility is that there are chemical inhomogeneities within the Galaxy and that the Sun formed from material richer than average in heavy elements.

      It is, in fact, clear that the Galaxy is not chemically homogeneous at the present time. Observations of radio recombination lines show that there is a gradient, or variation, in the temperature of nebulae throughout the Galaxy. This gradient almost surely implies a variation in the principal coolant, oxygen, and presumably in other heavy elements as well. The oxygen abundance is perhaps twice the solar value at one-third the Sun's distance from the galactic centre, and down to two-thirds of the solar abundance at the most distant points for which reliable determinations can be made (about 1.5 times the Sun's distance, or 45,000 light-years). These differences in heavy-element content reflect varying amounts of nucleosynthesis by massive stars. Similar gradients are found in other galaxies.

      The composition of nebulae in other galaxies can be determined by direct optical observations of emission lines. This method is not practical throughout the Milky Way because of the obscuration of dust. The Large Magellanic Cloud has compositions that are uniformly about one-half those of the Orion Nebula for oxygen, neon, argon, and sulfur and are one-quarter those of Orion for carbon and nitrogen. It appears that the first group of elements must be manufactured together, presumably in massive stars, and ejected together into the interstellar gas that is currently observable. Stars of a different mass (probably lower) must produce carbon and nitrogen. Planetary nebulae also suggest the same scenario.

      The abundance of helium in nebulae has received considerable attention because the helium content of the oldest objects provides clues to the origin of the universe. The value cited above for the Orion Nebula is in agreement with the predictions of the big-bang model, the prevailing cosmological theory according to which the universe began with an enormous explosion involving rapid expansion from a highly compressed primordial state. In order to determine the precise nature of this so-called big bang, a more precise estimate of helium abundance is needed than can be presently derived from nebulae.

Planetary nebulae (planetary nebula)
 There are about 20,000 objects called planetary nebulae in the Galaxy, each representing gas expelled relatively recently from a central star very late in its evolution. By contrast, diffuse and dark nebulae are clouds of gas from which young stars form. Because of the obscuration of dust in the Galaxy, only about 1,500 planetary nebulae have been cataloged.

Forms and structure
 Compared to the diffuse nebulae, planetary nebulae are small objects, having a radius, typically, of one light-year and containing a mass of gas equivalent to about 0.3 solar mass. One of the largest known planetary nebulae, the Helix Nebula (NGC 7293) in the constellation Aquarius, subtends an angle of about 20′ of arc—two-thirds the angular size of the full Moon. Planetary nebulae are considerably denser than most diffuse nebulae, typically containing 1,000–10,000 atoms per cubic centimetre within their dense regions, and have a surface brightness 1,000 times as large. Many are so far away that they appear stellar when photographed directly, but the conspicuous examples have an angular size up to 20′ of arc across, 10″–30″ being usual. Those that show a bright disk have much more regular forms than the chaotic diffuse nebulae, but there are still usually some brightness fluctuations over the disk. The planetaries generally have regular, sharp outer boundaries; often they have a relatively regular inner boundary as well, giving them an appearance like a ring. Many have two lobes of bright material, resembling arcs of a circle, connected by a bridge—somewhat resembling the letter Z.

 Most planetaries show a central star, called the nucleus, which provides the ultraviolet radiation required for ionizing (ionization) the gas in the ring or shell surrounding it. Those stars are among the hottest known and are in a state of comparatively rapid evolution.

      As with diffuse nebulae, the overall structural regularity conceals large-scale fluctuations in density. High-resolution photographs of a planetary nebula usually reveal tiny knots and filaments down to the resolution limit of the photograph. The spectrum of the planetary nebula is basically the same as that of the diffuse nebula; it contains bright lines from hydrogen and helium recombinations and the forbidden lines (defined above) of other ions (ion). In general, the spectra of planetaries differ from those of diffuse nebulae in that they show much higher degrees of ionization. In some planetaries most of the helium is doubly ionized, and appreciable amounts of five-times-ionized oxygen and argon and four-times-ionized neon exist. In diffuse nebulae helium is mainly once ionized and neon and argon only once or twice. This difference in the states of the atoms results from the temperature of the planetary nucleus (up to about 150,000 K), which is much higher than that of the exciting star of the diffuse nebula (less than 60,000 K for an O star, the hottest). One of the conspicuous features of planetaries is that the high stages of ionization are found close to the central star. The rare heavy ions, rather than hydrogen, absorb the photons of several hundred volts' energy. Beyond a certain distance from the central star, all the photons of energy sufficient to ionize a given species of ion have been absorbed, and that species therefore cannot exist farther out. Detailed theoretical calculations have rather successfully predicted the spectra of the best-observed nebulae.

      The spectra of planetary nebulae reveal another interesting fact: they are expanding from the central star at 24 to 56 km/s. The gravitational pull of the star is quite small at the distance of the shell from the star, so the shell will continue its expansion until it finally merges with the interstellar gas around it. The expansion is proportional to the distance from the central star, consistent with the entire mass of gas having been ejected at one brief period from the star in some sort of instability.

The distances of planetary nebulae
 Not a single planetary nebula is close enough to the Sun to allow a direct determination of its distance; the nearest, reasonably bright one (Helix Nebula) is about 300 light-years away. It is necessary, therefore, to use indirect methods to find the distances. One method is to assume a nebular mass, estimate the true rate of emission of one of the hydrogen lines from the nebula (from the gas density and temperature), and compare this with the observed brightness. From the intrinsic and apparent brightness the distance can be found. One planetary, NGC 246, has a normal G star as a companion, whose distance can be estimated if it is similar in intrinsic brightness to nearby G stars of similar type. Once the distance is known, other quantities (such as size and mass) can easily be determined. Unfortunately, different methods of estimating distances give results differing by factors of more than two. A more reliable way of estimating distance is to consider the planetaries found in the nearest external galaxies, such as the Andromeda Galaxy or the Magellanic Clouds. The distances to the galaxies are known from their other stars, such as the Cepheid variables. Even the most luminous planetary nebulae are very faint in other galaxies, and so only the few brightest and atypically massive nebulae can be studied.

      From the best available average distance determination, the true size of any nebula can be found, within the limits of error, from its angular size; typically, planetary nebulae are found to be a few tenths of a light-year in radius. If this distance is divided by the expansion speed, the age of the nebula since ejection is obtained, with values up to roughly 30,000 years. After this time, the expanding gas becomes indistinguishable from the interstellar medium.

Chemical composition
      The chemical composition of planetaries can be found from their spectra in the way discussed above for diffuse nebulae. Planetary nebulae definitely show signs of chemical enrichment from elements produced by nuclear processing within the central star. Some are carbon-rich, with twice as much carbon as oxygen, while the opposite is true for the Sun. Others are overabundant in nitrogen; the most luminous ones, observed in external galaxies, are conspicuous examples. Helium is modestly enhanced, up to a factor of two over the solar value. There is one object that contains almost no hydrogen; it is as if the gas had been ejected from the object at the very end of the nuclear-burning process. Planetary nebulae also show a clear indication of the general heavy-element abundance gradient in the Galaxy, presumably a reflection of the original composition of the stars that gave rise to the present nebulae.

      Some, but not all, planetary nebulae contain internal dust (interplanetary dust particle). In general, this dust cannot be seen directly but can be detected from the infrared radiation it emits after being heated by nebular and stellar radiation. The presence of dust implies that planetary nebulae are even richer in heavy elements than gas-phase abundance studies suggest.

      Among nebulae so far discovered, two are particularly deviant in chemical composition: one is in the globular cluster M15 and the other in the halo (galactic halo) (tenuous outer regions) of the Galaxy. Both have very low heavy-element content (down from normal by factors of about 50) but normal helium. Both objects are very old, suggesting that the primeval gas in the Galaxy had a low heavy-element content but an almost normal amount of helium. The origin of helium in the Galaxy was probably the initial explosion of the universe itself.

Positions in the Galaxy
      One of the best indicators of the average age of astronomical objects is their position and motion in the Galaxy. The youngest are in the spiral arms, near the gas from which they have formed; the oldest are not concentrated in the plane of the Galaxy, nor are they found within the spiral arms. By these criteria, the planetaries reveal themselves to be rather middle-aged; they are moderately but not strongly concentrated in the plane; rather, they are concentrated toward the galactic centre, as the older objects are. Their motions in the Galaxy follow elliptical paths, whereas circular orbits are characteristic of younger stars. They belong to the type of distribution often called a “disk population,” to distinguish them from the Population II (very old) and Population I (young) objects proposed by the German-American astronomer Walter Baade. It is likely that there is a wide variation in the ages of planetaries and that some are very young objects.

Evolution of planetary nebulae
      A description of the evolution of a planetary nebula begins before the ejection of the nebula itself. As will be discussed below, the central star is a red giant before the ejection. In such a phase it experiences a rapid loss of mass, up to 0.01 Earth mass per day, in the form of a comparatively slowly expanding stellar wind. At this stage the red giant might be heavily obscured by dust, which forms from the heavy elements in the wind. Eventually the nature of both the star and its wind changes. The star becomes hotter because its hot core is exposed by the loss of the overlying atmosphere. The inner gas is ionized by radiation from the hot star. The ionization zone moves steadily outward through the slowly moving material of what was formerly the stellar wind. The expansion speed of the gas is typically 30 km/s. Nebulae in this stage are bright but have starlike images as seen from the Earth, because they are too small to show a disk. The gas is at a relatively high density—about one million atoms per cubic centimetre—but becomes more dilute as the gas expands. During this stage the nebula is surrounded by neutral hydrogen. It appears to expand faster than the individual atoms of gas in it are moving; the ionized shell is “eating into” the neutral material as the density falls.

      The middle stage of evolution occurs when the density has dropped to the point at which the entire mass of gas is ionized. After this stage is reached, some of the ultraviolet radiation escapes into space, and the expansion of the nebula is caused entirely by the motion of the gas. Most planetaries are now in this middle stage. Finally the central star becomes less luminous and can no longer provide enough ultraviolet radiation to keep even the dilute nebula ionized. Once again the outer regions of the nebula become neutral and therefore invisible. Eventually the gas is mingled with the general interstellar gas.

      A curious feature of several planetaries is that faint rings surrounding the bright, inner nebula can be observed. Possibly the outer rings are the remnants of a previous shell ejected earlier by the star. Alternately, they might be at the outer edge of the expanding neutral (and invisible) gas, which is ramming into quiescent interstellar gas originally present when the planetary ejection occurred. There would be a strong shock wave at this interface, and the shocked matter would be heated to several thousand kelvins, causing the emission. The matter would quickly radiate its heat and become cold and invisible again. The key to the choice of those alternatives would be the spectrum of the faint ring, if it could be determined.

The central stars
 The central stars are known from their spectra to be very hot. A common type of spectrum has very broad emission lines of carbon or nitrogen, as well as of ionized helium, superimposed upon a bluish continuum. These spectra are indistinguishable from those from the very bright rare stars known as Wolf-Rayet stars (Wolf-Rayet star), but the planetary nuclei are about 100 times fainter than true Wolf-Rayet objects. The stars appear to be losing some mass at the present time, though evidently not enough to contribute appreciably to the shell.

      The presence of the nebula allows a fairly precise determination of the central star's evolution. The temperature of the star can be estimated from the nebula from the amounts of emission of ionized helium and hydrogen by a method devised by the Dutch astronomer H. Zanstra. The amount of ionized-helium radiation is determined by the number of photons of more than 54 volts' energy, while hydrogen is ionized by photons in excess of 13.6 volts. The relative numbers of photons in the two groups depend strongly on temperature, since the spectrum shifts dramatically to higher energies as the temperature of the star increases. Hence, the temperature can be found from the observed strengths of the hydrogen and helium lines. The rate of evolution of the stars can be determined from the sizes of their nebulae, as the time since ejection of the shell is the radius of the nebula divided by the expansion rate. The energy output, or luminosity, of the central star can be estimated from the brightness of the nebula, because the nebula is converting the star's invisible ultraviolet radiation (which contains the greater part of the star's luminous energy) into visible radiation.

      The resulting theoretical description of the star's evolution is quite interesting. While there seem to be real differences in stars at a given stage, the trends are quite clear. The central stars in the youngest planetary nebulae are about as hot as the massive O and B stars—35,000–40,000 K—but roughly 10 times fainter. They have half the diameter of the Sun but are 1,000 times as luminous. As the nebula expands, the star increases its brightness and temperature, but its radius decreases steadily. It reaches a maximum energy output, when it is roughly 10,000 times as luminous as the Sun, about 5,000 years after the initial expansion. This is an amazingly small fraction of the star's age of several times 109 years; it represents a period equivalent to about half an hour in a human life. From this point on the star becomes fainter, but for some time the temperature continues to increase while the shrinkage of the star continues. At its hottest the star is perhaps 200,000 K, almost five times hotter than the hottest of most of the stars. It then cools and after about 10,000 years becomes a very dense white dwarf star, scarcely larger than the Earth but with a density of thousands of kilograms per cubic centimetre. From this point it cools very slowly, becoming redder and fainter indefinitely.

      While there is not yet a very detailed theoretical picture of this contraction, a few results have emerged rather clearly: (1) white dwarf stars must obtain nearly all of their energy from the contraction noted above, not from nuclear sources; therefore, (2) they must contain practically no hydrogen or helium, except perhaps in a very thin shell on their surfaces. These conditions would have to be met for the evolution to take place so quickly.

      The absence of hydrogen in the star's interior is quite surprising; the planetary nebulae are all found to have a normal hydrogen abundance of about 1,000 times as many hydrogen atoms as heavy elements, such as oxygen. Thus, the mechanism of expulsion of the envelope must be very efficient at ejecting the hydrogen-rich outer layers of a star while leaving heavy-element-rich material behind.

The nature of the progenitor stars
      The progenitor must have mass not much in excess of a solar mass because of the distribution of the planetaries in the Galaxy. Very massive stars are young and more closely confined than are nebulae to the galactic plane. Also, the mass of the nebula is roughly 0.3 solar mass, and the mass of a typical white dwarf (the final state of the central star) is roughly 0.7 solar mass. Next, the expansion velocity of the nebula is probably comparable with the velocity of escape from its progenitor, which implies that the progenitor was a red giant star, large and cool, completely unlike the small, hot, blue, nuclear star remaining after the ejection. Likely candidates are members of the class of long-period variable stars, which have about the right size and mass and are known to be unstable. Symbiotic stars (i.e., stars with characteristics of both cool giants and very hot stars) also are candidates. Novas, stars that brighten temporarily while ejecting a shell explosively, are definitely not candidates; the nova shell is expanding at hundreds of kilometres per second.

      The cause of the ejection is the outward force of radiation on the outer layers of red giant stars. The ejection is triggered by a rapid variation in the nuclear luminosity in the interior of the giant, caused by instability in the helium-burning shell. The ejection takes place during more than one phase of the giant's evolution. Nitrogen-rich nebulae develop during an early episode when convection inside the star carries nitrogen, produced from carbon in a series of nuclear reactions (i.e., the carbon-nitrogen cycle of hydrogen burning), to the surface. A later ejection takes place with an enrichment of both nitrogen and helium, which also is produced by hydrogen burning. A still later phase occurs when convection carries carbon, the product of helium burning, to the surface.

Supernova remnants
  The supernova phenomenon is a spectacular explosion in which a star ejects most of its mass in a violently expanding cloud of debris that soon becomes a nebula. At the brightest phase of the explosion, the expanding cloud radiates as much energy in a single day as the Sun has done in the past three million years. Such explosions occur roughly every 50 years within a large galaxy. They have been observed less frequently in the Milky Way Galaxy because most of them have been hidden by the obscuring clouds of dust. Galactic supernovas were observed in 1006 in Lupus, in 1054 in Taurus, in 1572 in Cassiopeia (Tycho's nova, named after Tycho Brahe, its observer), and finally in 1604 in Serpens, called Kepler's Nova. The stars became bright enough to be visible in the daytime. The only naked-eye supernova to occur since 1604 was Supernova 1987A in the Large Magellanic Cloud (the galaxy nearest to the Milky Way system), but it was visible only from the Southern Hemisphere. On Feb. 23, 1987, a blue supergiant star brightened to gradually become third magnitude, easily visible at night, and has subsequently been followed in every wavelength band available to scientists. The spectrum showed hydrogen lines expanding at 12,000 km/s, followed by a long period of slow decline that is expected to continue for years.

      Supernova remnants evolve through four stages as they expand. At first, they expand so violently that they simply sweep all older interstellar material before them, acting as if they were expanding into a vacuum. The shocked gas, heated to millions of kelvins by the explosion, does not radiate its energy very well and is readily visible only in X-ray emissions. This stage typically lasts several hundred years, after which time the shell has a radius of about 10 light-years. As the expansion occurs, little energy is lost but the temperature falls because the same energy is spread into an ever-larger volume. The lower temperature favours more emission, and during the second phase the supernova remnant radiates its energy at the outermost, coolest layers. This phase can last thousands of years. The third stage occurs after the shell has swept up a mass of interstellar material that is comparable to or greater than its own; the expansion is then slowed substantially. The dense material, mostly interstellar at its outer edge, radiates away its remaining energy for hundreds of thousands of years. The final phase is reached when the pressure within the supernova remnant becomes comparable to the pressure of the interstellar medium outside the remnant, so that the remnant loses its distinct identity. In the later stages of expansion, the magnetic field of the galaxy is important in determining the motions of the weakly expanding gas. Even after the bulk of the material has merged with the local interstellar medium there might be regions of very hot gas remaining from the supernova explosion that produce soft X rays (i.e., those of a few hundred volts) observable locally and also confine the clumps of the diffuse ionized gas and diffuse nebulae.

      The recent galactic supernovas observed are in the first phases of the evolution suggested above. At the sites of Kepler's and Tycho's novas (Tycho's Nova), there exist heavy obscuring clouds, and the optical objects remaining are now inconspicuous knots of glowing gas. Near Tycho's nova, in Cassiopeia, there are similar optically insignificant wisps that appear to be remnants of yet another supernova explosion. To a radio telescope, however, the situation is spectacularly different: the Cassiopeia remnant is the strongest radio source in the entire sky. Study of this remnant, called Cassiopeia A, reveals that a supernova explosion occurred there in approximately 1680, missed by observers because of the obscuring dust.

 At the site of the 1054 supernova is one of the most remarkable objects in the sky, the Crab Nebula. Photographed in colour, it is revealed as a beautiful red, lacy network of long and sinuous, glowing hydrogen filaments surrounding a bluish, structureless region whose light is strongly polarized. The filaments emit the spectrum characteristic of a diffuse nebula. The gas is expanding at 1,100 km/s—but still slower than the 6,400 km/s in the shells of new supernovas in other galaxies. The bluish, amorphous inner region of the Crab Nebula is even more remarkable. The glow is produced by the so-called synchrotron radiation—i.e., from electrons spiraling about a magnetic field at almost the speed of light. This radiation is dramatically different from what it would be if it were emitted from electrons moving at low speeds: it becomes (1) strongly concentrated in the forward direction, (2) spread out over a broad range of frequencies but with the average frequency increasing with the electron's energy, and (3) highly polarized. Electrons of many different energies produce radiation in a large range of wavelengths: radio, infrared, optical, and ultraviolet. Even X rays (X-ray source) are emitted copiously by the Crab, which is the second-brightest X-ray source in the sky after Scorpius X-1 (an X-ray binary star), indicating that the electrons must have energies in the cosmic-ray range. After almost 1,000 years, the nebula is still losing 100,000 times as much energy per second as the Sun.

      On the basis of this huge outpouring of energy, it is easy to calculate how long the nebula can shine without a new supply of energy. The electrons emitting the X rays should decay, or drop to lower energies, in about 30 years—far less than the age of the nebula. The source of energy within the Crab Nebula supplying energy to the electrons that emit the X rays was discovered in 1969 to be a pulsar designated NP 0532, the remains of the stellar body that exploded to form the nebula.

      In the case of the Crab Nebula, the pulsar has been found to flash optically as well as at radio wavelengths, blinking on and off with the same period, owing to its rotation: 0.033099324 second (on June 28, 1969). This period is slowly increasing (at the rate of 0.0012 second per century), which implies the pulsar is slowing down and thereby losing its energy to the nebula. The corresponding rate of energy loss is about equal to the nebula's rate of energy loss, convincing evidence that a tiny, extremely dense pulsar can supply the energy to the nebula, which is about four light-years across.

      The Crab Nebula is unique in being a young supernova remnant and relatively free from obscuration, while Tycho's and Kepler's supernovas are conspicuous radio sources, apparently radiating by synchrotron emission; in neither case has a detectable pulsar been found. The failure to detect pulsars at the sites of Kepler's and Tycho's novas, or Cassiopeia A, does not mean that they are not there; the radiation from the pulsar is probably strongly beamed, and the Earth may not be located in the direction to which the pulsars are sending their energy.

      The Crab Nebula is still in a stage of violent expansion. The gas will continue to expand and the pulsar lose energy until the nebula enters a second stage: one in which the pulsar can no longer put much energy into the nebula, and the nebula expands like a hot bubble of gas into the cold surrounding interstellar gas. As the hot gas rams into the cold much faster than the speed of sound, a strong shock wave results. The cold gas is heated to several million kelvins as it is hit by the shock wave and then cools by radiating forbidden and recombination lines. Almost all the mass inside the shocked “bubble” will consist of this material swept up by the expanding shock wave.

  As a general mechanism this picture seems well corroborated. The best-observed old supernova remnant is the Cygnus Loop (or the Veil Nebula), a beautiful, filamentary object roughly in the form of a circular arc in Cygnus. Its patchiness is striking: the loop consists of a series of wisps rather than a continuous cloud of gas (see image—>). The most likely interpretation of this patchiness is that the interstellar medium into which the shock wave is propagating contains small clouds of denser material; many lines of reasoning from other evidence lead to the same result. The present speed of the filaments is about 100 km/s; the approximate age of the Cygnus Loop is 50,000 years.

Other supernova remnants
 Many other such old supernova remnants in the Galaxy are recognized, because they are also sources of radio emission from synchrotron radiation. Optically, in general, they show a wispy appearance similar to that of the Cygnus Loop.

      Supernova remnants modify the interstellar medium in two important ways: (1) They leave behind the elements that have been made in the presupernova star by a long process of heating and thermonuclear processing. Some very heavy elements are made in the extremely high temperatures right after the explosion. These elements are ejected into the supernova nebula and are finally merged with the interstellar medium. In this way the heavy-element content of the interstellar medium is increased. There is observational confirmation of this; both the Crab Nebula and the remnant of Tycho's supernova seem heavy-element rich. (2) Supernova remnants provide kinetic energy to the surrounding gas by their violent expansion, heating and accelerating the interstellar gas clouds. In addition, they provide large quantities of very high energy cosmic rays (cosmic ray) by accelerating them in the expanding shock wave. These cosmic rays, as well as the X rays emitted by the remnant, heat the interstellar gas. Some supernovas explode within the hot phase of the interstellar medium that has remained from previous supernovas and from stellar winds. In this case, the violently expanding shock reheats the tenuous gas that fills most of the volume of space in a galaxy but makes up very little of its mass. Thus, supernovas serve to maintain a hot component of gas, which is revealed only in X rays.

Diffuse ionized gas
      A recently recognized major component of interstellar gas has been discovered by means of its Hα emission line at 6563 angstroms, produced by a hydrogen atom dropping from the third to the second energy level. The American astronomer Ronald Reynolds and his collaborators have used a Fabry-Pérot interferometer to map and study this line and a few others (N+, S+, O++). The Hα line is produced by an H+ ion recombining with a free electron, exactly as in diffuse nebulae. Each Hα photon requires about two recombinations of an H+ ion, and each of these recombinations must be balanced by an ionization. The Hα line is faint, but it is seen in every direction, and the total amount of energy required to produce it is amazingly large. The net energy requirement of the Hα is about 15 percent of the luminosity of all O and B stars that provide ionization for diffuse nebulae. This energy output is about equal to the total power provided by all known supernovas, but the latter radiate the large bulk of their energy in nonionizing radiation or in providing modest kinetic energies to large amounts of gas near the end of their expansion phase. Hence, supernovas cannot power the diffuse ionized gas. Other potential sources of ionization, such as very hot, highly evolved stars of various types, such as Wolf-Rayet stars, white dwarf stars, or neutron stars, all fall far short. Cosmic rays are very energetic particles rather inefficiently accelerated by supernovas and so are less able to provide the required power than supernovas themselves. Only normal galactic stars of types O and B, similar to those that ionize diffuse nebulae, are viable candidates.

      Unlike H II regions, the diffuse ionized gas is found far from the galactic plane as well as close to it. Pulsars (pulsar) occasionally occur at large distances from the galactic plane. The pulses of radio radiation emitted by these objects are slowed in their passage through the interstellar medium by electrons from the ionized hydrogen. The delays of the pulses of various radio frequencies show that the electrons (and hence H+) extend about 3,000 light-years above and below the plane, indicating a much greater distance than the 300-light-year thickness of the neutral hydrogen, diffuse nebulae, or O and B stars in general. In the diffuse ionized gas, the comparatively low stages of ionization of the common elements (O+, N+, S+) are much more abundant relative to higher stages (O++, N++, S++) as compared to typical diffuse nebulae. Such an effect is caused by the extremely low density of the diffuse ionized gas; in this case, even hot stars fail to produce high stages of ionization. Thus, it seems possible to explain the peculiar ionization of the diffuse ionized gas with ionization powered by O and B stars. However, there is a major problem associated with this explanation: the ionizing radiation must be able to reach the diffuse ionized gas from the stars, which are typically surrounded by neutral hydrogen clouds that should prevent any ionizing radiation from penetrating through them. Apparently the O and B stars are able to ionize passages through the clouds enveloping them, and a substantial part of the ionizing radiation can escape into the regions far from the galactic plane. No other explanation seems plausible.

Additional Reading
An excellent general reference work on all topics in astronomy, including all forms of nebulae, is Stephen P. Maran (ed.), The Astronomy and Astrophysics Encyclopedia (1992), with articles written for the technically trained person who is not familiar with astronomy. Works appropriate for the lay reader include Lyman Spitzer, Jr., Searching Between the Stars (1982), written by one of the most authoritative figures in the field; J.E. Dyson and D.A. Williams, Physics of the Interstellar Medium (1980); and numerous articles written at a semitechnical level in the journals Scientific American (monthly); Mercury (bimonthly), published by the Astronomical Society of the Pacific; Astronomy (monthly); and Sky and Telescope (monthly). Readers with a thorough knowledge of physics may wish to consult Lyman Spitzer, Jr., Physical Processes in the Interstellar Medium (1978), very comprehensive and very terse; Donald E. Osterbrock, Astrophysics of Gaseous Nebulae and Active Galactic Nuclei (1989); and Stuart R. Pottasch, Planetary Nebulae: a Study of Late Stages of Stellar Evolution (1984).John S. Mathis

* * *


Universalium. 2010.

Игры ⚽ Поможем написать курсовую

Look at other dictionaries:

  • Nebula-H — Жанр EBM, electro industrial Годы С 2001 Страна …   Википедия

  • NEBULA — inter metcora infimae regionis notissima, memoratur Plinio l. 2. c. 60. quam densam in bello maximi momenti esse, discimus cx Livio l. 22. c. 4. ubi de Flaminio Consule ab Annibale circumvento, Eo magis, ait, Romanis subita atque improvisa res… …   Hofmann J. Lexicon universale

  • Nébula — est une super vilaine fictive créée par Marvel Comics, apparue pour la première fois dans Avengers #257, en 1985. Origine La luphomoïde Nébula, redoutable pirate de l espace, s empara du Sanctuaire III, un énorme vaisseau appartenant à Thanos. Ce …   Wikipédia en Français

  • Nebula — (Mehrzahl Nebulae) ist Lateinisch für Nebel (althochdeutsch nebul), abgeleitet aus dem griechischen nephele (»Wolke«). Nebula kann stehen für: verschiedene Objekte in der Astronomie: Nebel (Astronomie) die als Nebula Device bekannte Grafik Engine …   Deutsch Wikipedia

  • Nebula — Neb u*la (n[e^]b [ u]*l[.a]), n.; pl. {Nebul[ae]} (n[e^]b [ u]*l[=e]). [L., mist, cloud; akin to Gr. nefe lh, ne fos, cloud, mist, G. nebel mist, OHG. nebul, D. nevel, Skr. nabhas cloud, mist. Cf. {Nebule}.] 1. (Astron.) A faint, cloudlike, self… …   The Collaborative International Dictionary of English

  • nebula — / nɛbula/ s.f. [dal lat. nebŭla ]. 1. (ant., lett.) [fenomeno atmosferico dovuto a condensazione di vapore acqueo: una n. di colore di fuoco (Dante)] ▶◀ nebbia, nube, nuvola. 2. (astron.) [oggetto celeste avente l aspetto di nube diffusa,… …   Enciclopedia Italiana

  • nebula — (n.) early 15c., nebule a cloud, mist, from L. nebula mist, vapor, fog, smoke, exhalation, figuratively darkness, obscurity, from PIE *nebh cloud (Cf. Skt. nabhas vapor, cloud, mists, fog, sky; Gk. nephele, nephos cloud; Ger. nebel fog; O.E …   Etymology dictionary

  • nébula — s. f. Névoa.   ‣ Etimologia: latim nebula, ae …   Dicionário da Língua Portuguesa

  • Nebŭla — Nebŭla, 1) (lat.), Nebel; 2) Verdunkelung der Hornhaut in Gestalt eines leichten Nebels. Daher Nebulist, 1) Wolkenmaler, Luftbildner; 2) flüchtiger Zeichner; Nebulistisch, nebelig, wolkig, nebelhaft; nebulistische Zeichnung, eine solche, welche… …   Pierer's Universal-Lexikon

  • Nebŭla — (lat.), Nebel; Nebulist, Wolkenmaler, Nebler; nebulistisch, nebelhaft …   Meyers Großes Konversations-Lexikon

  • Nebula — Nebula, lat., Nebel; Nebulist, Wolkenmaler; nebulistisch zeichnen, nur schwache Umrisse hervortreten lassen; nebulös, nebelig, umwölkt, gräulich; Nebulosität, Umwölkung, finsteres Wesen …   Herders Conversations-Lexikon

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”