telescope

telescope
/tel"euh skohp'/, n., adj., v., telescoped, telescoping.
n.
1. an optical instrument for making distant objects appear larger and therefore nearer. One of the two principal forms (refracting telescope) consists essentially of an objective lens set into one end of a tube and an adjustable eyepiece or combination of lenses set into the other end of a tube that slides into the first and through which the enlarged object is viewed directly; the other form (reflecting telescope) has a concave mirror that gathers light from the object and focuses it into an adjustable eyepiece or combination of lenses through which the reflection of the object is enlarged and viewed. Cf. radio telescope.
2. (cap.) Astron. the constellation Telescopium.
adj.
3. consisting of parts that fit and slide one within another.
v.t.
4. to force together, one into another, or force into something else, in the manner of the sliding tubes of a jointed telescope.
5. to shorten or condense; compress: to telescope the events of five hundred years into one history lecture.
v.i.
6. to slide together, or into something else, in the manner of the tubes of a jointed telescope.
7. to be driven one into another, as railroad cars in a collision.
8. to be or become shortened or condensed.
[1610-20; TELE-1 + -SCOPE; r. telescopium ( < NL; see -IUM) and telescopio ( < It)]

* * *

Device that collects light from and magnifies images of distant objects, undoubtedly the most important investigative tool in astronomy.

The first telescopes focused visible light by refraction through lenses; later instruments used reflection from curved mirrors (see optics). Their invention is traditionally credited to Hans Lippershey (1570?–1619?), who adapted A. van Leeuwenhoek's use of lenses in microscopes. Among the earliest telescopes were Galilean telescopes, modeled after the simple instruments built by Galileo, who was the first to use telescopes to study celestial bodies. In 1611 Johannes Kepler proposed an improved version that became the basis for modern refracting instruments. The reflecting telescope came into its own after William Herschel (see Herschel family) used one to discover the planet Uranus in 1781. Since the 1930s radio telescopes have been used to detect and form images from radio waves emitted by celestial objects. More recently, telescopes have been designed to observe objects and phenomena in other parts of the electromagnetic spectrum (see gamma-ray astronomy; infrared astronomy; ultraviolet astronomy; X-ray astronomy). Spaceflight has allowed telescopes to be launched into Earth orbit to avoid the light-scattering and light-absorbing effects of the atmosphere (e.g., the Hubble Space Telescope). See also binoculars; observatory.

* * *

Introduction

      device used to form magnified images of distant objects.

      The telescope is undoubtedly the most important investigative tool in astronomy. It provides a means of collecting and analyzing radiation from celestial objects, even those in the far reaches of the universe.

       Galileo revolutionized astronomy when he applied the telescope (Galilean telescope) to the study of extraterrestrial bodies in the early 17th century. Until then, magnification instruments had never been used for this purpose. Since Galileo's pioneering work, increasingly more powerful optical telescopes have been developed, as has a wide array of instruments capable of detecting and measuring invisible forms of radiation, such as radio, X-ray, and gamma-ray telescopes. Observational capability has been further enhanced by the invention of various kinds of auxiliary instruments (e.g., the camera, spectrograph, and charge-coupled device) and by the use of electronic computers, rockets, and spacecraft in conjunction with telescope systems. These developments have contributed dramatically to advances in scientific knowledge about the solar system, the Milky Way Galaxy, and the universe as a whole.

General considerations
 Today, the telescope is used to explore every region of the electromagnetic spectrum from the shortest wavelengths (gamma rays) to the longest (radio waves; see Figure 1—>). The wavelengths of the spectrum are measured in three different units: angstroms (angstrom) (Å), micrometres (micrometre) (μ), and metres (metre) (m). Each of these units is customarily used for specific wavelength ranges, as shown in the figure. For example, the wavelengths for gamma rays and X rays are given in angstroms, those for infrared rays in micrometres, and those for intermediate radio waves in metres. (Centimetres are often used for short radio waves [microwaves] and kilometres for long radio waves.)

      Astronomical observations were restricted to visible wavelengths until the 1930s, when Karl Jansky and Grote Reber of the United States opened the radio “window.” Since the 1960s the use of Earth-orbiting telescope systems has enabled astronomers to make observations in all other spectral regions as well.

Optical telescopes

Refracting telescopes
  Commonly known as refractors, telescopes of this kind are used to examine the visible-light region of the electromagnetic spectrum. Typical uses include viewing the Moon, other objects of the solar system such as Jupiter and Mars, and double stars. The name refractor is derived from the term refraction, which is the bending of light when it passes from one medium to another of different density—e.g., from air to glass. The glass is referred to as a lens and may have one or more components. The physical shape of the components may be convex, concave, or plane-parallel. Figure 2—> illustrates the principle of refraction and the term focal length. The focus is the point, or plane, at which light rays from infinity converge after passing through a lens and traveling a distance of one focal length. In a refractor, the first lens through which light from a celestial object passes is called the objective lens. It should be noted that the light will be inverted at the focal plane. A second lens, referred to as the eyepiece lens, is placed behind the focal plane and enables the observer to view the enlarged, or magnified, image. Thus, the simplest form of refractor consists of an objective and an eyepiece, as illustrated in Figure 3—>.

      The diameter of the objective is referred to as the aperture; it typically ranges from a few centimetres for small spotting telescopes up to one metre for the largest refractor in existence. The objective, as well as the eyepiece, may have several components. Small spotting telescopes may contain an extra lens behind the eyepiece to erect the image so that it does not appear upside-down. When an object is viewed with a refractor, the image may not appear sharply defined, or it may even have a predominant colour in it. Such distortions, or aberrations, are sometimes introduced when the lens is polished into its design shape. The major kind of distortion in a refractor is chromatic aberration, which is the failure of the differently coloured light rays to come to a common focus. Chromatic aberration can be minimized by adding components to the objective. In lens-design technology, the coefficients of expansion of different kinds of glass are carefully matched to minimize the aberrations that result from temperature changes of the telescope at night.

      Eyepieces, which are used with both refractors and reflectors (see below Reflecting telescopes (telescope)), have a wide variety of applications and provide observers with the ability to select the magnification of their instruments. The magnification, sometimes referred to as magnifying power, is determined by dividing the focal length of the objective by the focal length of the eyepiece. For example, if the objective has a focal length of 254 centimetres (100 inches) and the eyepiece has a focal length of 2.54 centimetres (1 inch), then the magnification will be 100. Large magnifications are very useful for observing the Moon and the planets; however, since stars appear as point sources owing to their great distances, magnification provides no additional advantage when viewing them. Another important factor that one must take into consideration when attempting to view at high magnification is the stability of the telescope mounting. Any vibration in the mounting will also be magnified and may severely reduce the quality of the observed image. Thus, great care is usually taken to provide a stable platform for the telescope. This problem should not be associated with that of atmospheric seeing, which may introduce a disturbance to the image due to fluctuating air currents in the path of the light from a celestial or terrestrial object. Generally, most of the seeing disturbance arises in the first 30 metres of air above the telescope. Large telescopes are frequently installed on mountain peaks in order to get above the seeing disturbances.

Light gathering and resolution
      The most important of all the powers of an optical telescope is its light-gathering power. This capacity is strictly a function of the diameter of the clear objective—that is, the aperture—of the telescope. Comparisons of different-sized apertures for their light-gathering power are calculated by the ratio of their diameters squared; for example, a 25-centimetre objective will collect four times the light of a 12.5-centimetre objective [(25 × 25) ÷ (12.5 × 12.5)] = 4. The advantage of collecting more light with a larger-aperture telescope is that one can observe fainter stars, nebulas, and very distant galaxies.

      Resolving power is another important feature of a telescope. This is the ability of the instrument to distinguish clearly between two points whose angular separation is less than the smallest angle that the observer's eye can resolve. The resolving power of a telescope can be calculated by the formula

      resolving power = 11.25 seconds of arc/d,

      where d is the diameter of the objective expressed in centimetres. Thus, a 25-centimetre-diameter objective has a theoretical resolution of 0.45 second of arc and a 250-centimetre telescope has one of 0.045 second of arc. An important application of resolving power is in the observation of visual binary stars. Here, one star is routinely observed as it revolves around a second star. Many observatories conduct extensive visual binary observing programs and publish catalogs of their observational results. One of the major contributors in this field is the United States Naval Observatory in Washington, D.C.

      Most refractors currently in use at observatories have equatorial mountings. (The mounting describes the orientation of the physical bearings and structure that permits a telescope to be pointed at a celestial object for viewing.) In the equatorial mounting, the polar axis of the telescope is constructed parallel to the Earth's axis. The polar axis supports the declination axis of the instrument. declination is measured on the celestial sky north or south from the celestial equator. The declination axis makes it possible for the telescope to be pointed at various declination angles as the instrument is rotated about the polar axis with respect to right ascension. Right ascension is measured along the celestial equator from the vernal equinox (i.e., the position on the celestial sphere where the Sun crosses the celestial equator from south to north on the first day of spring). Declination and right ascension are the two coordinates that define a celestial object on the celestial sphere. Declination is analogous to latitude, and right ascension is analogous to longitude. Graduated dials are mounted on the axis to permit the observer to point the telescope precisely. To track an object, the telescope's polar axis is driven smoothly by an electric motor at a sidereal rate—namely, at a rate equal to the rate of rotation of the Earth with respect to the stars. Thus, one can track or observe with a telescope for long periods of time if the sidereal rate of the motor is very accurate. High-accuracy, motor-driven systems have become readily available with the rapid advancement of quartz-clock technology. Most major observatories now rely on either quartz or atomic clocks to provide accurate sidereal time for observations as well as to drive telescopes at an extremely uniform rate.

       Some important ground-based optical telescopesA notable example of a refracting telescope is the 66-centimetre refractor of the U.S. Naval Observatory (United States Naval Observatory). This instrument was used by the astronomer Asaph Hall to discover the two moons of Mars, Phobos and Deimos, in 1877. Today, the telescope is used primarily for observing double stars. The 91-centimetre refractor at Lick Observatory on Mount Hamilton, Calif., U.S., and the one-metre instrument at Yerkes Observatory in Williams Bay, Wis., U.S., are the largest refracting systems currently in operation (Table 1 (Some important ground-based optical telescopes)).

      Another type of refracting telescope is the astrograph, which usually has an objective diameter of approximately 20 centimetres. The astrograph has a photographic plateholder mounted in the focal plane of the objective so that photographs of the celestial sphere can be taken. The photographs are usually taken on glass plates. The principal application of the astrograph is to determine the positions of a large number of faint stars. These positions are then published in catalogs such as the AGK3 and serve as reference points for deep-space photography.

Reflecting telescopes
  Reflectors are used not only to examine the visible region of the electromagnetic spectrum but also to explore both the shorter- and longer-wavelength regions adjacent to it (i.e., the ultraviolet and the infrared). The name of this type of instrument is derived from the fact that the primary mirror reflects the light back to a focus instead of refracting it. The primary mirror usually has a concave spherical or parabolic shape, and, as it reflects the light, it inverts the image at the focal plane. Figure 4—> illustrates the principle of a concave reflecting mirror. The formulas for resolving power, magnifying power, and light-gathering power, as discussed for refractors, apply to reflectors as well.

      The primary mirror is located at the lower end of the telescope tube in a reflector and has its front surface coated with an extremely thin film of metal, such as aluminum. The back of the mirror is usually made of glass, although other materials have been used from time to time. Pyrex (trademark) was the principal glass of choice for many of the older large telescopes, but new technology has led to the development and widespread use of a number of glasses with very low coefficients of expansion. A low coefficient of expansion means that the shape of the mirror will not change significantly as the temperature of the telescope changes during the night. Since the back of the mirror serves only to provide the desired form and physical support, it does not have to meet the high optical quality standards required for a lens.

 Reflecting telescopes have a number of other advantages over refractors. They are not subject to chromatic aberration because reflected light does not disperse according to wavelength. Also, the telescope tube of a reflector is shorter than that of a refractor of the same diameter, which reduces the cost of the tube. Consequently, the dome for housing a reflector is smaller and more economical to construct. So far only the primary mirror for the reflector has been discussed. In examining Figure 4—>, one might wonder about the location of the eyepiece. The primary mirror reflects the light of the celestial object to the prime focus near the upper end of the tube. Obviously, if an observer put his eye there to observe with a modest-sized reflector, he would block out the light from the primary mirror with his head. Isaac Newton (Newton, Sir Isaac) placed a small plane mirror at an angle of 45° inside the prime focus and thereby brought the focus to the side of the telescope tube. The amount of light lost by this procedure is very small when compared to the total light-gathering power of the primary mirror. The Newtonian reflector is popular among amateur telescope makers.

 A contemporary of Newton, N. Cassegrain of France, invented another type of reflector. Called the Cassegrainian telescope (Cassegrain reflector), this instrument employs a small convex mirror to reflect the light back through a small hole in the primary mirror to a focus located behind the primary. Figure 5—> illustrates a typical Cassegrain reflector. Some large telescopes of this kind do not have a hole in the primary mirror but use a small plane mirror in front of the primary to reflect the light outside the main tube and provide another place for observation. The Cassegrain design usually permits short tubes relative to their mirror diameter.

      One more variety of reflector was invented by another of Newton's contemporaries, the Scottish astronomer James Gregory (Gregory, James). Gregory placed a concave secondary mirror outside the prime focus to reflect the light back through a hole in the primary mirror. Notable is the fact that the Gregorian design was adopted for the Earth-orbiting space observatory, the Solar Maximum Mission (SMM), launched in 1980.

      Most large reflecting telescopes that are currently in use have a cage at their prime focus that permits the observer to sit inside the telescope tube while operating the instrument. The five-metre reflector at Palomar Observatory, near San Diego, Calif., is so equipped. While most reflectors have equatorial mounts similar to refractors, the world's largest reflector, the six-metre instrument at the Special Astrophysical Observatory in Zelenchukskaya, Russia, has an altitude-azimuth mounting. The significance of the latter design is that the telescope must be moved in both altitude and azimuth as it tracks a celestial object. Equatorial mountings, by contrast, require motion in only one coordinate while tracking, since the declination coordinate is constant. Reflectors, like refractors, usually have small guide telescopes mounted parallel to their main optical axis to facilitate locating the desired object. These guide telescopes have low magnification and a wide field of view, the latter being a desirable attribute for finding stars or other remote cosmic objects.

      The parabolic shape of a primary mirror has a basic failing in that it produces a narrow field of view. This can be a problem when one wishes to observe extended celestial objects. To overcome this difficulty, most large reflectors now have a modified Cassegrain design. The central area of the primary mirror has its shape deepened from that of a paraboloid, and the secondary mirror is configured to compensate for the altered primary. The result is the Ritchey-Chrétien design, which has a curved rather than a flat focus. Obviously, the photographic medium must be curved to collect high-quality images across the curved focal plane. The one-metre telescope of the U.S. Naval Observatory in Flagstaff, Ariz., was one of the early examples of this design.

 The above-mentioned Ritchey-Chrétien design has a good field of view of about 1°. For some astronomical applications, however, photographing larger areas of the sky is mandatory. In 1930 Bernhard Schmidt (Schmidt, Bernhard Voldemar), an optician at the Hamburg Observatory in Bergedorf, Ger., designed a catadioptric telescope that satisfied the requirement of photographing larger celestial areas. A catadioptric telescope design incorporates the best features of both the refractor and reflector—i.e., it has both reflective and refractive optics. The Schmidt telescope has a spherically shaped primary mirror. Since parallel light rays that are reflected by the centre of a spherical mirror are focused farther away than those reflected from the outer regions, Schmidt introduced a thin lens (called the correcting plate) at the radius of curvature of the primary mirror. Since this correcting plate is very thin, it introduces little chromatic aberration. The resulting focal plane has a field of view several degrees in diameter. Figure 6—> illustrates a typical Schmidt design.

      The National Geographic Society–Palomar Observatory Sky Survey made use of a 1.2-metre Schmidt telescope to photograph the northern sky in the red and blue regions of the visible spectrum. The survey produced 900 pairs of photographic plates (about 7° by 7° each) taken from 1949 to 1956. Schmidt telescopes of the European Southern Observatory in Chile and of the observatory at Siding Spring Mountain in Australia have photographed the remaining part of the sky that cannot be observed from Palomar Mountain. (The survey undertaken at the latter included photographs in the infrared as well as in the red and blue spectral regions.)

Multimirror telescopes
      The main reason astronomers build larger telescopes is to increase light-gathering power so that they can see deeper into the universe. Unfortunately, the cost of constructing larger single-mirror telescopes increases rapidly—approximately with the cube of the diameter of the aperture. Thus, in order to achieve the goal of increasing light-gathering power while keeping costs down, it has become necessary to explore new, more economical and nontraditional telescope designs.

 The American-built Multiple Mirror Telescope (MMT Observatory) (MMT), located at the Whipple Observatory in Arizona, represents such an effort. The MMT has six 1.8-metre paraboloid mirrors mounted in a single framework; the light from all the mirrors is concentrated into a single focus. The mirrors, under computer control, are automatically aligned at regular intervals during an observing tour. A 10-metre multimirror telescope was installed on Mauna Kea on the island of Hawaii in 1992, and a second telescope was completed in 1996. Each of the Keck telescopes comprise 36 contiguous, adjustable mirror segments, all under computer control. Even larger multimirror instruments are currently being planned by American and European astronomers.

Special types of optical telescopes
Solar telescopes
      Either a refractor or reflector may be used for visual observations of solar features, such as sunspots or solar prominences. The main precaution that needs to be taken is to reduce the intensity of the image so that the observer's eye will not be damaged. Generally, this is done with a tinted eyepiece. Special solar telescopes have been constructed, however, for investigations of the Sun that require the use of such ancillary instruments as spectroheliographs and coronagraphs. These telescopes are mounted in towers and have very long focus objectives. Typical examples of tower solar telescopes are found at the Mount Wilson Observatory in California and the McMath-Hulbert Observatory in Michigan. The long focus objective produces a very good scale factor, which in turn makes it possible to look at individual wavelengths of the solar electromagnetic spectrum in great detail. A tower telescope has an equatorially mounted plane mirror at its summit to direct the sunlight into the telescope objective. This plane mirror is called a coelostat. Bernard Lyot (Lyot, Bernard) constructed another type of solar telescope in 1930 at Pic du Midi Observatory in France. This instrument was specifically designed for photographing the Sun's corona (the outer layer), which up to that time had been successfully photographed only during solar eclipses. The coronagraph, as this special telescope is called, must be located at a high altitude to be effective. The high altitude is required to reduce the scattered sunlight, which would reduce the quality of the photograph. The High Altitude Observatory in Colorado and the Sacramento Peak Observatory in New Mexico have coronagraphs.

Earth-orbiting space telescopes
 While astronomers continue to seek new technological breakthroughs with which to build larger ground-based telescopes, it is readily apparent that the only solution to some scientific problems is to make observations from above the Earth's atmosphere. A series of Orbiting Astronomical Observatories (OAOs (Orbiting Astronomical Observatory)) has been launched by the National Aeronautics and Space Administration (NASA); the OAO launched in 1972 (later named Copernicus) had an 81-centimetre telescope on board. The most sophisticated observational system (satellite observatory) placed in Earth orbit so far is the Hubble Space Telescope (HST; see photograph—>). Launched in 1990, the HST is essentially an optical-ultraviolet telescope with a 2.4-metre primary mirror. It has been designed to enable astronomers to see into a volume of space 300 to 400 times larger than that permitted by other systems. At the same time, the HST is not impeded by any of the problems caused by the atmosphere. It is equipped with five principal scientific instruments: (1) a wide-field and planetary camera, (2) a faint-object spectrograph, (3) a high-resolution spectrograph, (4) a high-speed photometer, and (5) a faint-object camera. The HST was launched into orbit from the U.S. Space Shuttle at an altitude of more than 570 kilometres above the Earth. Shortly after its deployment in Earth orbit, HST project scientists found that a manufacturing error affecting the shape of the telescope's primary mirror severely impaired the instrument's focusing capability. The flawed mirror causes spherical aberration, which limits the ability of the HST to distinguish between cosmic objects that lie close together and to image distant galaxies and quasars. Project scientists devised measures that enabled them to compensate in part for the defective mirror and correct the imaging problem.

Astronomical transit instruments
      These small but extremely important telescopes play a vital role in mapping the celestial sphere. Without the transit instrument's very accurate determination of stellar and planetary positions, the larger deep-space telescopes would not be able to find their desired celestial object.

      Astronomical transit instruments are usually refractors with apertures of 15 to 20 centimetres. (Ole Rømer (Rømer, Ole), a Danish astronomer, is credited with having invented this type of telescope system.) The main optical axis of the instrument is aligned on a north-south line such that its motion is restricted to the plane of the meridian of the observer. The observer's meridian is a great circle on the celestial sphere that passes through the north and south points of the horizon as well as through the zenith of the observer. Restricting the telescope to motion only in the meridian provides an added degree of stability, but it requires the observer to wait for the celestial object to rotate across his meridian. The latter process is referred to as transiting (transit) the meridian, from which the name of the telescope is derived. There are various types of transit instruments, as, for example, the transit circle telescope, the vertical circle telescope, and the horizontal meridian circle telescope. The transit circle determines the right ascension of celestial objects, while the vertical circle measures only their declinations. Transit circles and horizontal meridian circles measure both right ascension and declination at the same time. The final output data of all transit instruments are included in star or planetary catalogs.

 One of the most accurate astronomical transit instruments in the world is the U.S. Naval Observatory's 15-centimetre transit circle telescope (see photograph—>). Other notable examples of this class of telescopes include the transit circle of the National Astronomical Observatory in Tokyo, the meridian circle of the Bordeaux Observatory in France, and the automatic meridian circle of the Roque de los Muchachos Observatory in the Canary Islands.

Astrolabes (astrolabe)
      Another special type of telescopic instrument is the modern version of the astrolabe. Known as a prismatic astrolabe, it too is used for making precise determinations of the positions of stars and planets. It may sometimes be used inversely to determine the latitude and longitude of the observer, assuming the star positions are accurately known. The aperture of a prismatic astrolabe is small, usually only 8 to 10 centimetres. A small pool of mercury and a refracting prism make up the other principal parts of the instrument. An image reflected off the mercury is observed along with a direct image to give the necessary position data. The most notable example of this type of instrument is the French-constructed Danjon astrolabe. During the 1970s, however, the Chinese introduced various innovations that resulted in a more accurate and automatic kind of astrolabe, which is now in use at the Peking Observatory.

B.L. Klock

Radio telescopes (radio telescope)
  Radio telescopes are used to study naturally occurring radio emissions from stars (star), galaxies (galaxy), quasars (quasar), and other astronomical (astronomy) objects between wavelengths of about 10 metres (30 megahertz [MHz]) and 1 mm (300 gigahertz [GHz]).

Principles of operation
      Radio telescopes vary widely, but they all have two basic components: (1) a large radio antenna and (2) a sensitive radiometer, or radio receiver. The sensitivity of a radio telescope—i.e., the ability to measure weak sources of radio emission—depends both on the area and efficiency of the antenna and on the sensitivity of the radio receiver used to amplify and to detect the signals. For broadband continuum emission over a range of wavelengths, the sensitivity also depends on the bandwidth of the receiver. Because cosmic radio sources are extremely weak, radio telescopes are usually very large, up to hundreds of metres across, and use the most sensitive radio receivers available. Moreover, weak cosmic signals can be easily masked by terrestrial radio interference, and great effort is taken to protect radio telescopes from man-made emissions.

 The most familiar type of radio telescope is the radio reflector consisting of a parabolic antenna, which operates in the same manner as a television satellite dish to focus the incoming radiation onto a small antenna called the feed, a term that originated with antennas used for radar transmissions (see figure—>). This type of telescope is also known as the dish, or filled-aperture, telescope. In a radio telescope the feed is typically a waveguide horn and transfers the incoming signal to the sensitive radio receiver. Solid-state amplifiers (amplifier) that are cooled to very low temperatures to reduce significantly their internal noise are used to obtain the best possible sensitivity.

      In some radio telescopes the parabolic (parabola) surface is equatorially mounted, with one axis parallel to the rotation axis of Earth. Equatorial mounts are attractive because they allow the telescope to follow a position in the sky as the Earth rotates by moving the antenna about a single axis parallel to the Earth's axis of rotation. But equatorially mounted radio telescopes are difficult and expensive to build. In most modern radio telescopes, a digital computer is used to drive the telescope about the azimuth and elevation axes to follow the motion of a radio source across the sky.

      In the simplest form of radio telescope, the receiver is placed directly at the focal point of the parabolic reflector, and the detected signal is carried by cable along the feed support structure to a point near the ground where it can be recorded and analyzed. However, it is difficult in this type of system to access the instrumentation for maintenance and repair, and weight restrictions limit the size and number of individual receivers that can be installed on the telescope. More often, a secondary reflector is placed in front of (Cassegrain focus) or behind (Gregorian focus) the focal point of the paraboloid to focus the radiation to a point near the vertex, or centre, of the main reflector. Multiple feeds and receivers may be located at the vertex where there is more room, where weight restrictions are less stringent, and where access for maintenance and repair is more straightforward. Secondary focus systems also have the advantage that both the primary and secondary reflecting surfaces may be carefully shaped so as to improve the gain over that of a simple parabolic antenna.

      Earlier radio telescopes used a symmetric tripod or quadrapod structure to hold the feed or secondary reflector, but such an arrangement blocks some of the incoming radiation, and the reflection of signals from the support legs back into the receiver distorts the response. In newer designs, the feed or secondary reflector is placed off the central axis and does not block the incoming signal. Off-axis radio telescopes are thus more sensitive and less affected by interference reflected from the support structure into the feed.

      The performance of a radio telescope is limited by various factors. The accuracy of a reflecting surface may depart from the ideal shape because of manufacturing irregularities. Wind load can exert force on the telescope. Thermal deformations cause differential expansion and contraction. As the antenna is pointed to different parts of the sky, deflections occur due to changes in gravitational forces. Departures from a perfect parabolic surface become important when they are a few percent or more of the wavelength of operation. Since small structures can be built with greater precision than larger ones, radio telescopes designed for operation at millimetre wavelengths are typically only a few tens of metres across, whereas those designed for operation at centimetre wavelengths range up to 300 metres (1,000 feet) in diameter. For operation at relatively long metre wavelengths where the reflecting surface need not have an accuracy better than a few centimetres, it becomes practical to build very large fixed structures in which the reflecting surface can be made of simple “chicken wire” fencing or even parallel rows of wires.

      Traditionally, the effect of gravity has been minimized by designing the movable structure to be as stiff as possible in order to reduce the deflections resulting from gravity. A more effective technique, based on the principle of homology, allows the structure to deform under the force of gravity, and the cross section and weight of each member of the movable structure are chosen to cause the gravitational forces to deform the reflecting structure into a new paraboloid with a slightly different focal point. It is then necessary only to move the feed or secondary reflector to maintain optimum performance. Homologous designs have become possible only since the development of computer-aided structural simulations known as the finite element method (solids, mechanics of).

      Some radio telescopes, particularly those designed for operation at very short wavelengths, are placed in protective enclosures called radomes that can nearly eliminate the effect of both wind loading and temperature differences throughout the structure. Special materials that exhibit very low absorption and reflection of radio waves have been developed for such structures, but the cost of enclosing a large antenna in a suitable temperature-controlled radome may be almost as much as the cost of the movable antenna itself.

      The cost of constructing an antenna with a very large aperture can be greatly reduced by fixing the structure to the ground and moving either the feed or the secondary reflector to steer the beam in the sky. However, for parabolic reflecting surfaces, the beam can be steered in this way over only a limited range of angle without introducing aberration and a loss of signal strength.

      Radio telescopes are used to measure broad- bandwidth continuum radiation as well as narrow-bandwidth spectroscopic features due to atomic and molecular lines found in the radio spectrum of astronomical objects. In early radio telescopes, spectroscopic observations were made by tuning a receiver across a sufficiently large frequency range to cover the various frequencies of interest. Because the spectrometer had a narrow frequency range, this procedure was extremely time-consuming and greatly restricted observations. Modern radio telescopes observe simultaneously at a large number of frequencies by dividing the signals up into as many as several thousand separate frequency channels that can range over a much larger total bandwidth of tens to hundreds of megahertz.

      The most straightforward type of radio spectrometer employs a large number of filters, each tuned to a separate frequency and followed by a separate detector that combines the signal from the various filters to produce a multichannel, or multifrequency, receiver. Alternatively, a single broad-bandwidth signal may be converted into digital form and analyzed by the mathematical process of autocorrelation and Fourier transforms (Fourier transform) (see below). In order to detect faint signals, the receiver output is often averaged over periods of up to several hours to reduce the effect of noise generated by thermal radiation in the receiver.

Radio interferometry (radio interferometer) and aperture synthesis
      The angular resolution, or ability of a radio telescope to distinguish fine detail in the sky, depends on the wavelength of observations divided by the size of the instrument. Yet even the largest antennas, when used at their shortest operating wavelength, have an angular resolution of only a few arc seconds, which is about 10 times poorer than the resolution of ground-based optical telescopes. Because radio telescopes operate at much longer wavelengths than do optical telescopes, radio telescopes need to be much larger than optical telescopes to achieve the same angular resolution.

      At radio wavelengths, the distortions introduced by the atmosphere are less important than at optical wavelengths, and so the theoretical angular resolution of a radio telescope can in practice be achieved even for the largest dimensions. Also, because radio signals are easy to distribute over large distances without distortion, it is possible to build radio telescopes of essentially unlimited dimensions. In fact, the history of radio astronomy has been one of solving engineering problems to construct radio telescopes of continually increasing angular resolution.

      The high angular resolution of radio telescopes is achieved by using the principles of interferometry to synthesize a very large effective aperture from a number of antennas. In a simple two-antenna radio interferometer, the signals from an unresolved, or “point,” source alternately arrive in phase (constructive interference) and out of phase (destructive interference) as Earth rotates and causes a change in the difference in path from the radio source to the two elements of the interferometer. This produces interference fringes in a manner similar to that in an optical interferometer. If the radio source has finite angular size, then the difference in path length to the elements of the interferometer varies across the source. The measured interference fringes from each interferometer pair thus depend on the detailed nature of the radio “brightness” distribution in the sky.

      Each interferometer pair measures one “Fourier component” of the brightness distribution of the radio source. Work by Sir Martin Ryle (Ryle, Sir Martin) and his colleagues in the1950s and '60s showed that movable antenna elements combined with the rotation of Earth can sample a sufficient number of Fourier components with which to synthesize the effect of a large aperture and thereby reconstruct high-resolution images of the radio sky. The laborious computational task of doing Fourier transforms to obtain images from the interferometer data is accomplished with high-speed computers and the fast Fourier transform (FFT), a mathematical technique that is specially suited for computing discrete Fourier transforms. In recognition of his contributions to the development of the Fourier synthesis technique, more commonly known as aperture synthesis, or earth-rotation synthesis, Ryle was awarded the 1974 Nobel Prize for Physics.

      During the 1960s the Swedish physicist Jan Hogbom developed a technique called CLEAN, which is used to remove the spurious responses from a celestial radio image caused by the use of discrete, rather than continuous, spacings in deriving the radio image. Further developments, based on a technique introduced in the early 1950s by the British scientists Roger Jennison and Francis Graham Smith, led to the concept of self-calibration, which uses the observed source as its own calibrator in order to remove errors in a radio image due to uncertainties in the response of individual antennas as well as small errors introduced by the propagation of radio signals through the terrestrial atmosphere. In this way radio telescopes are able to achieve extraordinary angular resolution and image quality that are not possible in any other wavelength band.

Very long baseline interferometry
      In conventional interferometers and arrays, coaxial cable, waveguide, or even fibre-optic links are used to distribute a common local-oscillator reference signal to each antenna and also to return the received signal from an individual antenna to a central laboratory where it is correlated with the signals from other antennas. In cases in which antennas are spaced more than a few tens of kilometres apart, however, it becomes prohibitively expensive to employ real physical links to distribute the signals. Very high frequency (VHF) or ultrahigh frequency (UHF) radio links have been used, but the need for a large number of repeater stations makes this impractical for spacings greater than a few hundred kilometres.

      Interferometer systems of essentially unlimited element separation can be formed by using the technique of very long baseline interferometry (VLBI). In early VLBI systems the signals received at each element were recorded by broad-bandwidth videotape recorders located at each antenna. More recently, with the advent of inexpensive, reliable computer disk drives, the data are recorded on disks. The disks are then transported to a common location where they are replayed and the signals combined to form interference fringes. The successful operation of a VLBI system requires that the tape recordings be synchronized within a few millionths of a second and that the local oscillator reference signal be stable to better than one part in a trillion. Recorded data from just a few hours of observation typically contain about one trillion bits of information, which is roughly equivalent to storing the entire contents of a modest-sized library. Hydrogen maser frequency standards are used to give a timing accuracy of only a few billionths of a second and a frequency stability of one part in a million billion.

radar techniques
      Techniques analogous to those used in military and civilian radar applications are sometimes employed with radio telescopes to study the surface of planets (planet) and asteroids (asteroid) in the solar system. By measuring the spectrum and the time of flight of signals reflected from planetary surfaces, it is possible to examine topographical features with a linear resolution as good as 1 km, deduce rates of rotation, and determine with great accuracy the distance to the planets. Radio signals reflected from the planets are weak, and high-power radar transmitters are needed in order to obtain measurable signal detections. The time it takes for a radar signal to travel to Venus and back, even at the closest approach of the planet to Earth, is about five minutes. For Saturn, it is more than two hours.

Major applications of radio telescopes
      Radio telescopes permit astronomers to study many kinds of extraterrestrial radio sources. These astronomical objects emit radio waves by one of several processes, including (1) thermal radiation from solid bodies such as the planets, (2) thermal, or bremsstrahlung, radiation from hot gas in the interstellar medium, (3) synchrotron radiation from electrons moving at velocities near the speed of light in weak magnetic fields, (4) spectral line radiation from atomic or molecular transitions that occur in the interstellar medium or in the gaseous envelopes around stars, and (5) pulsed radiation resulting from the rapid rotation of neutron stars (neutron star) surrounded by an intense magnetic field and energetic electrons.

      Radio telescopes are used to measure the surface temperatures of all the planets, as well as some of the moons of Jupiter and Saturn. Radar measurements have revealed the rotation of Mercury, which was previously thought to keep the same side toward the Sun. Astronomers have also used radar observations to image features on the surface of Venus, which is completely obscured from visual scrutiny by the heavy cloud cover that permanently enshrouds the planet. Accurate measurements of the travel time of radar signals reflected from Venus when it is on the other side of the Sun from Earth have indicated that radio waves passing close to the Sun slow down owing to gravity and thereby provide a new independent test of Albert Einstein (Einstein, Albert)'s theory of general relativity (relativity).

      Broadband continuum emission throughout the radio-frequency spectrum is observed from a variety of stars (star) (especially binary, X-ray, and other active stars), from supernova remnants, and from magnetic fields and relativistic electrons in the interstellar medium. The discovery of pulsars (pulsar) (short for pulsating radio stars) in 1967 revealed the existence of rapidly rotating neutron stars throughout the Milky Way Galaxy and led to the first observation of the effect of gravitational radiation (gravitation).

      Using radio telescopes equipped with sensitive spectrometers, radio astronomers have discovered about 150 separate molecules, including familiar chemical compounds like water, formaldehyde, ammonia, methanol, ethyl alcohol, and carbon dioxide. The important spectral line of atomic hydrogen at 1,421.405 MHz (21 cm wavelength) is used to determine the motions of hydrogen clouds in the Milky Way Galaxy and other galaxies. This is done by measuring the change in the wavelength of the observed lines arising from the Doppler effect. It has been established from such measurements that the rotational velocities of the hydrogen clouds vary with distance from the galactic centre. The mass of a spiral galaxy can in turn be estimated using this velocity data. In this way radio telescopes show evidence for the presence of so-called dark matter by showing that the amount of starlight is insufficient to account for the large mass inferred from the rapid rotation curves.

      Radio telescopes have discovered powerful radio galaxies (galaxy) and quasars (quasar) far beyond the Milky Way Galaxy system. These cosmic objects have intense clouds of radio emission that extend hundreds of thousands of light-years away from a central energy source located in an active galactic nucleus (AGN), or quasar. Observations with high-resolution radio arrays show highly relativistic jets extending from an AGN to the radio lobes. (For more specific information about quasars and other extragalactic radio sources, see cosmos: Quasars and related objects (Cosmos) and galaxy: Quasars (galaxy).)

      Measurements made in 1965 by Arno Penzias (Penzias, Arno) and Robert Wilson (Wilson, Robert Woodrow) using an experimental communications antenna at 3 cm wavelength located at Bell Laboratories in Holmdel, N.J., detected the existence of a microwave cosmic background radiation with a temperature of 3 kelvins (K). This radiation, which comes from all parts of the sky, is thought to be the remaining radiation from the hot big bang, the primeval explosion from which the universe presumably originated 13.7 billion years ago. Satellite and ground-based radio telescopes have been used to measure the very small deviations from isotropy of the cosmic microwave background. This work has led to refined determination of the size, geometry, and age of the universe.

Important radio telescopes
Filled-aperture telescopes
       Some important radio telescopes Some important radio telescopesThe largest single radio telescope in the world is the 305-metre (1,000-foot) fixed spherical reflector operated by Cornell University at the Arecibo Observatory near Arecibo, P.R. The antenna has an enormous collecting area, but the beam can be moved through only a limited angle of about 20° from the zenith. It is used for planetary radar astronomy, as well as for studying pulsars and other galactic and extragalactic phenomena.

      The Russian RATAN-600 telescope (RATAN stands for Radio Astronomical Telescope of the Academy of Sciences), located near Zelenchukskaya in the Caucasus Mountains, has 895 reflecting panels, each 7.4 metres (24.3 feet) high, arranged in a ring 576 metres (1,890 feet) in diameter. Using long parabolic cylinders, standing reflectors, or dipole elements, researchers in Australia, France, India, Italy, Russia, and Ukraine have also built antennas with very large collecting areas.

      The largest fully steerable radio telescope in the world is the Robert C. Byrd Green Bank Telescope (GBT) located in Green Bank, W.Va. This 110-by-100-metre (360-by-330-foot) off-axis radio telescope was completed in 2000 and operates at wavelengths as short as a few millimetres. The moving structure, which weighs 7.3 million kg (16 million pounds), points to any direction in the sky with an accuracy of only a few arc seconds. The secondary reflector is held by an off-axis support structure to minimize radiation from the ground and unwanted reflections from support legs. Each of the 2,004 surface panels that make up the parabolic surface is held in place by computer-controlled actuators that keep the surface accurate to a few tenths of a millimetre. The GBT is located in the National Radio Quiet Zone, which offers unique protection for radio telescopes from local sources of man-made interference.

 Other large, fully steerable, filled-aperture radio telescopes include the Max Planck Institut für Radioastronomie 100-metre- (330-foot-) diameter antenna near Effelsberg, Ger.; the Australian Commonwealth Scientific and Industrial Research Organization (CSIRO) 64-metre (210-foot) dish near Parkes; and the 76-metre (250-foot) Lovell Telescope at Jodrell Bank (Jodrell Bank Observatory) in England. These filled-aperture radio telescopes are used for atomic and molecular spectroscopy over a wide range of frequency and for other galactic and extragalactic studies.

      Several smaller, more precise radio telescopes for observing at millimetre wavelength have been installed high atop mountains or other high elevations, where clear skies and high altitudes minimize absorption and distortion of the incoming signals by the terrestrial atmosphere. A 45-metre (148-foot) radio dish near Nobeyama, Japan, is used for observations at wavelengths as short as 3 mm (0.12 inch). The French-Spanish Institut de Radio Astronomie Millimetrique (IRAM) in Grenoble, France, operates a 30-metre (100-foot) antenna at an altitude of 2,850 metres (9,350 feet) on Pico Veleta in the Spanish Sierra Nevada for observations at wavelengths as short as 1 mm (0.04 inch). Several radio telescopes that operate at submillimetre wavelengths are located near the summit of Mauna Kea, Hawaii, at elevations above 4,000 metres (13,000 feet) and on Mount Graham near Tucson, Ariz. The largest of these, the James Clerk Maxwell Telescope at the Mauna Kea Observatory, has a diameter of 15 metres (49 feet).

Radio telescope arrays
  The world's most powerful radio telescope, in its combination of sensitivity, resolution, and versatility, is the U.S. Very Large Array (VLA) located on the Plains of San Agustin near Socorro, in central New Mexico. The VLA consists of 27 parabolic antennas, each measuring 25 metres (82 feet) in diameter. The total collecting area is equivalent to a single 130-metre (430-foot) antenna. However, the angular resolution is equivalent to a single antenna 36 km (22 miles) in diameter. Each element of the VLA can be moved by a transporter along a Y-shaped railroad track; it is possible to change the length of the arms between 600 metres (2,000 feet) and 21 km (13 miles) to vary the resolution. Each antenna is equipped with receivers that operate in eight different wavelength bands from approximately 7 mm (0.3 inch) to 4 metres (13 feet). When used at the shorter wavelength in the largest antenna configuration, the angular resolution of the VLA is better than one tenth of an arc second, or about the same as the Hubble Space Telescope at optical wavelengths. The VLA is operated by the U.S. National Radio Astronomy Observatory as a facility of the National Science Foundation and is used by nearly 1,500 astronomers each year for a wide variety of research programs devoted to the study of the solar system, the Milky Way Galaxy, radio stars, pulsars, atomic and molecular gas in the Milky Way Galaxy and in other galaxies, radio galaxies, quasars, and the radio afterglow of gamma-ray bursters (gamma-ray burster).

      In Europe the Netherlands Foundation for Research in Astronomy operates the Westerbork Synthesis Radio Telescope, which is an east-west array of 14 antennas, each 25 metres (82 feet) in diameter and extending over 2.7 km (1.7 miles). In Australia the Commonwealth Scientific and Industrial Research Organization maintains the six-element Australian Telescope Compact Array at Narrabri, N.S.W., for studies of the southern skies, including in particular the nearby Magellanic Clouds (Magellanic Cloud).

      Indian radio astronomers have built the Giant Metrewave Radio Telescope (GMRT) near Pune, India. The GMRT contains 30 antennas extending some 25 km (16 miles) in diameter. Each antenna element is 45 metres (148 feet) in diameter and is constructed using a novel, inexpensive system of wire trusses to replace the conventional steel beam backup structure of the parabolic surface. The GMRT operates at relatively long wavelengths between 20 cm (8 inches) and 6 metres (20 feet).

      The Multi-Element Radio Linked Interferometer Network (MERLIN), operated by the Nuffield Radio Astronomy Laboratories at Jodrell Bank, is being upgraded to use fibre-optic, instead of microwave radio, links to connect seven antennas separated by up to 217 km (135 miles) in the southern part of England. It is used primarily to study compact radio sources associated with quasars, AGN, and cosmic masers with a resolution of a few hundreths of an arc second.

      The Very Long Baseline Array (VLBA) consists of ten 25-metre (82-foot) dishes spread across the United States from the Virgin Islands to Hawaii. The VLBA operates at wavelengths from 3 mm (0.1 inch) to 1 metre (3 feet) and is used to study quasars, galactic nuclei, cosmic masers, pulsars, and radio stars with a resolution as good as 0.0001 arc second, or more than 100 times better than that of the Hubble Space Telescope. The 10 individual antenna elements of the VLBA do not have any direct connection; instead, signals are recorded on high-density computer disk drives that are then shipped to a special processing centre in New Mexico where they are replayed and the signals analyzed to form images. Precise timing between the elements is maintained by a hydrogen maser atomic clock located at each antenna site. The control and analysis centre for the VLBA is located in central New Mexico along with the VLA Operations centre, and the two instruments are sometimes used together to obtain increased sensitivity and angular resolution.

      In 1997, Japanese radio astronomers working at the Institute for Space Science near Tokyo launched an 8-metre (26-foot) dish, known as the VLBI Space Observatory Program (VSOP), in Earth orbit. Working with the VLBA and other ground-based radio telescopes, VSOP gave interferometer baselines up to 33,000 km (21,000 miles). (VSOP was also known as Highly Advanced Laboratory for Communication and Astronomy [HALCA].) In 2003 the VSOP lost its ability to point accurately, and the program ended.

      Interferometers and arrays are also used at millimetre and submillimetre wavelengths where they are used to study the formation of stars and galaxies with resolution better than can be obtained with simple filled-aperture antennas. The operation of arrays at millimetre and submillimetre wavelengths is very difficult and requires that the instrument be at very high and dry locations to minimize the phase distortions of signals as they propagate through the atmosphere. Some prominent millimetre interferometers and arrays are the Combined Array for Research in Millimeter-wave Astronomy (CARMA) near Big Pine, Calif., the IRAM Plateau de Bure facility in France, and the Japanese Nobeyama Radio Observatory. In 2003 the Harvard-Smithsonian Center for Astrophysics in collaboration with the Academia Sinica of Taiwan completed the Submillimeter Array (SMA) located near the summit of Mauna Kea, Hawaii, at an elevation of 4,080 metres (13,385 feet). This is an eight-element array of 6-metre (20-foot) dishes designed to work at wavelengths as short as 0.3 mm (0.01 inch). A major new international facility is under construction by the United States, Canada, Europe, and Japan in the Atacama Desert in northern Chile at an elevation of more than 5,000 metres (16,000 feet) and is expected to be completed by 2012. The Atacama Large Millimeter Array (ALMA) will consist of fifty 12-metre (39-foot) dishes operating at wavelengths as short as 0.3 mm (0.01 inch), as well as a more compact array of four 12-metre and sixteen 7-metre (23-foot) dishes.

Earth-orbiting radio telescopes
      Most radio waves pass relatively undistorted through Earth's atmosphere, and so there is little need to place radio telescopes in space. The exceptions are for observations at very long wavelengths that are distorted by Earth's ionosphere (ionosphere and magnetosphere), for observations at very short wavelengths that are affected by water vapour and oxygen in the atmosphere, and for precision observations at all wavelengths that might be affected by thermal radiation from the ground. The first radio astronomy satellite was the U.S.-British Ariel 2, launched in 1964, which studied long-wavelength radio noise from Earth's ionosphere and the Milky Way Galaxy. Ariel 2 was followed by two more satellites in the Ariel series and by the U.S. satellites Radio Astronomy Explorers 1 and 2, launched in 1968 and 1973, respectively.

      Subsequent radio astronomy satellites performed observations that were difficult to make from the ground or were enhanced by being made from space (such as VSOP mentioned above). The U.S. Submillimeter Wave Astronomy Satellite (SWAS) and the Swedish-Canadian-French-Finnish ODIN, launched in 1998 and 2001, respectively, observed at very short, submillimetre wavelengths. By observing the cosmic microwave background radiation left over from the big bang, the U.S. satellites Cosmic Background Explorer (launched in 1989) and the Wilkinson Microwave Anistropy Probe (launched in 2001) detected very small fluctuations in the background radiation corresponding to the early structures from which galaxies would be formed and accurately determined the age and composition of the universe.

Kenneth I. Kellermann

Other types of telescopes

Infrared telescopes
      Telescopic systems of this type do not really differ significantly from reflecting telescopes designed to observe in the visible region of the electromagnetic spectrum. The main difference is in the physical location of the infrared telescope, since infrared photons have lower energies than those of visible light. The infrared rays are readily absorbed by the water vapour in the Earth's atmosphere, and most of this water vapour is located at the lower atmospheric regions—i.e., near sea level. Earth-bound infrared telescopes have been successfully located on high mountaintops, as, for example, Mauna Kea in Hawaii. The other obvious placement of infrared instruments is in a satellite such as the Infrared Astronomical Satellite (IRAS), which mapped the celestial sky in the infrared in 1983. The Kuiper Airborne Observatory, operated by NASA, consists of a 0.9-metre telescope that is flown in a special airplane above the water vapour to collect infrared data. Much of the infrared data is collected with an electronic camera, since ordinary film is unable to register the low-energy photons.

      Another example of an infrared telescope is the United Kingdom Infrared Telescope (UKIRT), which has a 3.8-metre mirror made of Cer-Vit (trademark), a glass ceramic that has a very low coefficient of expansion. This instrument is configured in a Cassegrain design and employs a thin monolithic primary mirror with a lightweight support structure. This telescope is located at Mauna Kea Observatory. The 3-metre Infrared Telescope Facility (IRTF), also located at Mauna Kea, is sponsored by NASA and operated by the University of Hawaii.

Ultraviolet telescopes
      These telescopes are used to examine the shorter wavelengths of the electromagnetic spectrum immediately adjacent to the visible portion. Like the infrared telescopes, the ultraviolet systems also employ reflectors as their primary collectors. Ultraviolet radiation is composed of higher-energy photons than infrared radiation, which means that photographic techniques as well as electronic detectors can be used to collect astronomical data. The Earth's stratospheric ozone layer, however, blocks all wavelengths shorter than 3000 angstroms from reaching ground-based telescopes. As this ozone layer lies at an altitude of 20 to 40 kilometres, astronomers have to resort to rockets and satellites to make observations from above it. Since 1978 an orbiting observatory known as the International Ultraviolet Explorer (IUE) has studied celestial sources of ultraviolet radiation. The IUE telescope is equipped with a 45-centimetre mirror and records data electronically down to 1000 angstroms. The IUE is in a synchronous orbit (i.e., its period of revolution around the Earth is identical to the period of the planet's rotation) in view of NASA's Goddard Space Flight Center in Greenbelt, Md., and so data can be transmitted to the ground station at the end of each observing tour and examined immediately on a television monitor.

      Another Earth-orbiting spacecraft, the Extreme Ultraviolet (EUV) Explorer satellite, which is scheduled to be launched in the early 1990s, is designed to survey the sky in the extreme ultraviolet region between 400 and 900 angstroms. It has four telescopes with gold-plated mirrors, the design of which is critically dependent on the transmission properties of the filters used to define the EUV band passes. The combination of the mirrors and filters has been selected to maximize the telescope's sensitivity to detect faint EUV sources. Three of the telescopes have scanners that are pointed in the satellite's spin plane. The fourth telescope, the Deep Survey/Spectrometer Telescope, is directed in an anti-Sun direction. Its function is to conduct a photometric deep-sky survey in the ecliptic plane for part of the mission and then to collect spectroscopic observations in the final phase of the mission.

X-ray telescopes (X-ray telescope)
 The X-ray telescope is used to examine the shorter-wavelength region of the electromagnetic spectrum adjacent to the ultraviolet region. The design of this type of telescope must be radically different from that of a conventional reflector. Since X-ray photons have so much energy, they would pass right through the mirror of a standard reflector. X rays must be bounced off a mirror at a very low angle if they are to be captured. (This technique is referred to as grazing incidence.) For this reason, the mirrors in X-ray telescopes are mounted with their surfaces only slightly off a parallel line with the incoming X rays, as seen in Figure 8—>. Application of the grazing-incidence principle makes it possible to focus X rays from a cosmic object into an image that can be recorded electronically.

      NASA launched a series of three High-Energy Astronomy Observatories (HEAOs) during the late 1970s to explore cosmic X-ray sources. HEAO-1 mapped the X-ray sources with high sensitivity and high resolution. Some of the more interesting of these objects were studied in detail by HEAO-2 (named the Einstein Observatory). HEAO-3 was used primarily to investigate cosmic rays and gamma rays.

      The European X-ray Observatory Satellite (EXOSAT), developed by the European Space Agency (ESA), was capable of greater spectral resolution than the Einstein Observatory and was more sensitive to X-ray emissions at shorter wavelengths. EXOSAT remained in orbit from 1983 to 1986. A much larger X-ray astronomy satellite was launched in June 1990 as part of a cooperative program involving the United States, Germany, and the United Kingdom. This satellite, called the Röntgensatellit (ROSAT), has two parallel grazing-incidence telescopes. One of them, the X-ray telescope (XRT), bears many similarities to the equipment of the Einstein Observatory but has a larger geometric area and better mirror resolution. The other telescope, the extended ultraviolet wide-field camera, has an imaging detector much like the X-ray HRI. A positive sensitive proportional counter will make it possible to survey the sky at X-ray wavelengths for the purpose of producing a catalog of 100,000 sources with a positional accuracy of better than 30 arc seconds. A wide-field camera with a 5°-diameter field of view is also part of the ROSAT instrument package. It is designed to produce an extended ultraviolet survey with arc minute source positions in this wavelength region, making it the first instrument with such capability. The ROSAT mirrors are gold-coated and will permit detailed examination of the sky from 6 to 100 angstroms.

Gamma-ray telescopes
      These instruments require the use of grazing-incidence techniques similar to those employed with X-ray telescopes. Gamma rays are the shortest (about 0.1 angstrom or less) known waves in the electromagnetic spectrum. As mentioned above, HEAO-3 was developed to collect data from cosmic gamma-ray sources. NASA and collaborative international agencies have numerous ongoing and planned projects in the area of gamma-ray astronomy. The scientific objectives of the programs include determining the nature and physical parameters of high-energy (up to 10 gigaelectron volts) astrophysical systems. Examples of such systems include stellar coronas, white dwarfs, neutron stars, black holes, supernova remnants, clusters of galaxies, and diffuse gamma-ray background. In addition to satellite investigations of these cosmic high-energy sources, NASA has an extensive program that involves the design and development of gamma-ray telescope systems for deployment in high-altitude balloons. All mirrors in gamma-ray telescopes have gold coatings similar to those in X-ray telescope mirrors.

The development of the telescope and auxiliary instrumentation

Evolution of the optical telescope
       Galileo is credited with having developed telescopes (Galilean telescope) for astronomical observation in 1609. While the largest of his instruments was only about 120 centimetres long and had an objective diameter of 5 centimetres, it was equipped with an eyepiece that provided an upright (i.e., erect) image. Galileo used his modest instrument to explore such celestial phenomena as the valleys and mountains of the Moon, the phases of Venus, and the four largest Jovian satellites, which had never been systematically observed before.

      The reflecting telescope was developed in 1668 by Newton, though John Gregory had independently conceived of an alternative reflector design in 1663. Cassegrain introduced another variation of the reflector in 1672. Near the end of the century, others attempted to construct refractors as long as 61 metres, but these instruments were too awkward to be effective.

      The most significant contribution to the development of the telescope in the 18th century was that of Sir William Herschel (Herschel, Sir William). Herschel, whose interest in telescopes was kindled by a modest 5-centimetre Gregorian, persuaded the king of England to finance the construction of a reflector with a 12-metre focal length and a 120-centimetre mirror. Herschel is credited with having used this instrument to lay the observational groundwork for the concept of extragalactic “nebulas”—i.e., galaxies outside the Milky Way system.

      Reflectors continued to evolve during the 19th century with the work of William Parsons, 3rd Earl of Rosse (Rosse, William Parsons, 3rd earl of), and William Lassell (Lassell, William). In 1845 Lord Rosse constructed in Ireland a reflector with a 185-centimetre mirror and a focal length of about 16 metres. For 75 years this telescope ranked as the largest in the world and was used to explore thousands of nebulas and star clusters. Lassell built several reflectors, the largest of which was on Malta; this instrument had a 124-centimetre primary mirror and a focal length of more than 10 metres. His telescope had greater reflecting power than that of Rosse, and it enabled him to catalog 600 new nebulas as well as to discover several satellites of the outer planets—Triton (Neptune's largest moon), Hyperion (Saturn's 8th moon), and Ariel and Umbriel (two of Uranus' moons).

      Refractor telescopes, too, underwent development during the 18th and 19th centuries. The last significant one to be built was the 1-metre refractor at Yerkes Observatory. Installed in 1897, it remains the largest refracting system in the world. Its objective was designed and constructed by the optician Alvan Clark, while the mount was built by the firm of Warner & Swasey.

      The reflecting telescope has predominated in the 20th century. The rapid proliferation of larger and larger instruments of this type began with the installation of the 2.5-metre reflector at the Mount Wilson Observatory near Pasadena, Calif., U.S. The technology for mirrors underwent a major advance when the Corning Glass Works (in Steuben county, N.Y., U.S.) developed Pyrex. This borosilicate glass, which undergoes substantially less expansion than ordinary glass, was used in the 5-metre Hale reflector built in 1950 at the Palomar Observatory. Pyrex also was utilized in the main mirror of the world's largest telescope, the 6-metre reflector of the Special Astrophysical Observatory in Zelenchukskaya. In recent years, much better materials for mirrors have become available. Cer-Vit, for example, was used for the 4.2-metre William Herschel Telescope of the Roque de los Muchachos Observatory in the Canary Islands, and Zerodur (trademark) for the 3.5-metre reflector at the German-Spanish Astronomical Center in Calar Alto, Spain.

B.L. Klock

Development of the radio telescope
      Extraterrestrial radio emission was first reported in 1933 by Karl Jansky (Jansky, Karl), an engineer at the Bell Telephone Laboratories, while he was searching for the cause of shortwave interference. Jansky had mounted a directional radio antenna on a turntable so that he could point it at different parts of the sky to determine the direction of the interfering signals. He not only detected interference from distant thunderstorms but also located a source of radio “noise” from the centre of the Milky Way Galaxy. This first detection of cosmic radio waves received much attention from the public but only passing notice from the astronomical community.

      Grote Reber (Reber, Grote), a radio engineer and amateur radio operator, built a 9.5-metre parabolic reflector in his backyard in Wheaton, Ill., U.S., to continue Jansky's investigation of cosmic radio noise. In 1944 he published the first radio map of the sky. After World War II ended, the technology that had been developed for military radar was applied to astronomical research. Radio telescopes of increasing size and sophistication were built first in Australia and Great Britain and later in the United States and other countries (see above Radio telescopes: Important radio telescopes (telescope)).

Kenneth I. Kellermann

Advances in auxiliary instrumentation
      Almost as important as the telescope itself are the auxiliary instruments that the astronomer uses to exploit the light received at the focal plane. Examples of such instruments are the camera, spectrograph, photomultiplier tube, charge-coupled device (CCD), and charge injection device (CID). Each of these instrument types is discussed below.

Cameras (camera)
      John Draper of the United States photographed the Moon as early as 1840 by applying the daguerreotype process. The French physicists A.-H.-L. Fizeau (Fizeau, Armand-Hippolyte-Louis) and J.-B.-L. Foucault (Foucault, Jean) succeeded in making a photographic image of the Sun in 1845. Five years later, astronomers at Harvard Observatory took the first photographs of the stars.

      The use of photographic equipment in conjunction with telescopes has benefited astronomers greatly, giving them two distinct advantages: first, photographic images provide a permanent record of celestial phenomena and, second, photographic plates integrate the light from celestial sources over long periods of time and thereby permit astronomers to see much fainter objects than they would be able to observe visually. Typically, the camera's photographic plate (or film) is mounted in the focal plane of the telescope. The plate (or film) consists of glass or of a plastic material that is covered with a thin layer of a silver compound. The light striking the photographic medium causes the silver compound to undergo a chemical change. When processed, a negative image results—i.e., the brightest spots (the Moon and the stars, for example) appear as the darkest areas on the plate or the film.

Spectrographs
 Newton noted the interesting way in which a piece of glass can break up light into different bands of colour, but it was not until 1814 that the German physicist Joseph von Fraunhofer discovered the lines of the solar spectrum and laid the basis for spectroscopy. The spectrograph, as illustrated in Figure 9—>, consists of a slit, a collimator, a prism for dispersing the light, and a focusing lens. The collimator is an optical device that produces parallel rays from a focal plane source—i.e., gives the appearance that the source is located at an infinite distance. The spectrograph enables astronomers to analyze the chemical composition of planetary atmospheres, stars, nebulas, and other celestial objects. A bright line in the spectrum indicates the presence of a glowing gas radiating at a wavelength characteristic of the chemical element in the gas. A dark line in the spectrum usually means that a cooler gas has intervened and absorbed the lines of the element characteristic of the intervening material. The lines also may be displaced to either the red end or the blue end of the spectrum. This effect was first noted in 1842 by the Austrian physicist Christian Johann Doppler. When a light source is approaching, the lines are shifted toward the blue end of the spectrum, and when the source is receding, the lines are shifted toward its red end. This effect, known as the Doppler effect, permits astronomers to study the relative motions of celestial objects with respect to the Earth's motion.

      The slit of the spectrograph is placed at the focal plane of the telescope. The resulting spectrum may be recorded photographically or with some kind of electronic detector, such as a photomultiplier tube, CCD, or CID. If no recording device is used, then the optical device is technically referred to as a spectroscope.

Photomultiplier tubes (photomultiplier tube)
      The photomultiplier tube is an enhanced version of the photocell, which was first used by astronomers to record data electronically. The photocell contains a photosensitive surface that generates an electric current when struck by light from a celestial source. The photosensitive surface is positioned just behind the focus. A diaphragm of very small aperture is usually placed in the focal plane to eliminate as much of the background light of the sky as possible. A small lens is used to focus the focal plane image on the photosensitive surface, which, in the case of a photomultiplier tube, is referred to as the photocathode. In the photomultiplier tube a series of special sensitive plates are arranged geometrically to amplify or multiply the electron stream. Frequently, magnifications of a million are achieved by this process.

      The photomultiplier tube has a distinct advantage over the photographic plate. With the photographic plate the relationship between the brightness of the celestial source and its registration on the plate is not linear. In the case of the photomultiplier tube, however, the release of electrons in the tube is directly proportional to the intensity of light from the celestial source. This linear relationship is very useful for working over a wide range of brightness. A disadvantage of the photomultiplier tube is that only one object can be recorded at a time. The output from such a device is sent to a recorder or digital storage device to produce a permanent record.

Charge-coupled devices
      The charge-coupled device uses a light-sensitive material on a silicon chip to electronically detect photons in a way similar to the photomultiplier tube. The principal difference is that the chip also contains integrated microcircuitry required to transfer the detected signal along a row of discrete picture elements (or pixels) and thereby scan a celestial object or objects very rapidly. When individual pixels are arranged simply in a single row, the detector is referred to as a linear array. When the pixels are arranged in rows and columns, the assemblage is called a two-dimensional array.

      Pixels can be assembled in various sizes and shapes. The Hubble Space Telescope has a CCD detector with a 1,600 × 1,600 pixel array. Actually, there are four 800 × 800 pixel arrays mosaicked together. The sensitivity of a CCD is 100 times greater than a photographic plate and so has the ability to quickly scan objects such as planets, nebulas, and star clusters and record the desired data. Another feature of the CCD is that the detector material may be altered to provide more sensitivity at different wavelengths. Thus, some detectors are more sensitive in the blue region of the spectrum than in the red region.

      Today, most large observatories use CCDs to record data electronically. Another similar device, the charge injection device, is sometimes employed. The basic difference between the CID and the CCD is in the way the electric charge is transferred before it is recorded; however, the two devices may be used interchangeably as far as astronomical work is concerned.

Impact of technological developments
Computers (computer)
      Besides the telescope itself, the electronic computer has become the astronomer's most important tool. Indeed, the computer has revolutionized the use of the telescope to the point where the collection of observational data is now completely automated. The astronomer need only identify the object to be observed, and the rest is carried out by the computer and auxiliary electronic equipment.

      A telescope can be set to observe automatically by means of electronic sensors appropriately placed on the telescope axis. Precise quartz or atomic clocks send signals to the computer, which in turn activates the telescope sensors to collect data at the proper time. The computer not only makes possible more efficient use of telescope time but also permits a more detailed analysis of the data collected than could have been done manually. Data analysis that would have taken a lifetime or longer to complete with a mechanical calculator can now be done within hours or even minutes with a high-speed computer.

      Improved means of recording and storing computer data also have contributed to astronomical research. Optical disc data storage technology, such as the CD-ROM (compact disc read-only memory) or the WORM (write-once read-many) disc, has provided astronomers with the ability to store and retrieve vast amounts of telescopic and other astronomical data. A 12-centimetre CD-ROM, for example, may hold up to 600 megabytes of data—the equivalent of 20 nine-track magnetic tapes or 1,500 floppy discs. A 13-centimetre WORM disc typically holds about 300 to 400 megabytes of data.

Rockets (rocket) and spacecraft
      As noted earlier, the quest for new knowledge about the universe has led astronomers to study electromagnetic radiation other than just visible light. Such forms of radiation, however, are blocked for the most part by the Earth's atmosphere, and so their detection and analysis can only be achieved from above this gaseous envelope.

      During the late 1940s, single-stage sounding rockets were sent up to 160 kilometres or more to explore the upper layers of the atmosphere. From 1957, more sophisticated multistage rockets were launched as part of the International Geophysical Year; these rockets carried artificial satellites equipped with a variety of scientific instruments. Beginning in 1959, the Soviet Union and the United States, engaged in a “space race,” intensified their efforts and launched a series of unmanned probes to explore the Moon. Lunar exploration culminated with the first manned landing on the Moon by the U.S. Apollo 11 astronauts on July 20, 1969. Numerous other U.S. and Soviet spacecraft were sent to further study the lunar environment until the mid-1970s.

      Starting in the early 1960s both the United States and the Soviet Union launched a multitude of unmanned deep-space probes to learn more about the other planets and satellites of the solar system. Carrying television cameras, detectors, and an assortment of other instruments, these probes sent back impressive amounts of scientific data and close-up pictures. Among the most successful missions were those involving the Soviet Venera probes to Venus and the U.S. Viking 1 and 2 landings on Mars and Voyager 2 flybys of Jupiter, Saturn, Uranus, and Neptune. When the Voyager 2 probe flew past Neptune and its moons in August 1989, every known major planet had been explored by spacecraft. Many long-held views, particularly those about the outer planets, were altered by the findings of the Voyager probe. These findings included the discovery of several rings and six additional satellites around Neptune, all of which are undetectable to ground-based telescopes.

      Specially instrumented spacecraft have enabled astronomers to investigate other celestial phenomena as well. The Orbiting Solar Observatories (OSOs) and Solar Maximum Mission, Earth-orbiting U.S. satellites equipped with ultraviolet detector systems, have provided a means for studying solar activity. Another example is the Giotto probe of the European Space Agency, which enabled astronomers to obtain detailed photographs of the nucleus of Comet Halley during the 1986 passage of the comet.

B.L. Klock

Additional Reading
Techniques of observation with the help of telescopes are detailed in P. Clay Sherrod and Thomas L. Koed, A Complete Manual of Amateur Astronomy: Tools and Techniques for Astronomical Observations (1981), including a discussion of telescope setup; Robert T. Dixon, Dynamic Astronomy, 5th ed. (1989), a comprehensive, well-illustrated text; and Jay M. Pasachoff, Astronomy, from the Earth to the Universe, 3rd ed. (1987). A highly readable and profusely illustrated discussion of radio telescopes is given in Gerrit L. Verschuur, The Invisible Universe Revealed: The Story of Radio Astronomy, 2nd ed. (2006). More technical accounts are offered in Bernard F. Burke and Francis Graham-Smith, An Introduction to Radio Astronomy, 2nd ed. (2002); Gerrit L. Verschuur and Kenneth Kellermann, Galactic and Extragalactic Radio Astronomy, 2nd ed. (1988); and K. Rohlfs and T.L. Wilson, Tools of Radio Astronomy, 4th ed. (2004). Bernard Lovell, The Story of Jodrell Bank (1968), presents a dramatic story of the construction of the 76-metre Jodrell Bank Radio Telescope. Other modern types of telescopes are described in D.S. Finley et al., “Design of the Extreme Ultraviolet Explorer Long-Wavelength Grazing Incidence Telescope Optics,” Applied Optics 27(8):1476–80 (April 15, 1988); Adelaide Hewitt (ed.), Optical and Infrared Telescopes for the 1990s, 2 vol. (1980); and Joseph F. Baugher, The Space-Age Solar System (1988). Numerous articles on various types of telescopes are found in Sky and Telescope (monthly). Historical surveys tracing the development of the telescope from the earliest versions to those constructed in the late 1980s include Arthur Berry, A Short History of Astronomy (1898, reprinted 1961); Henry C. King, The History of the Telescope (1955, reprinted 1979); and Richard Learner, Astronomy Through the Telescope (1981).B.L. Klock Kenneth I. Kellermann

* * *


Universalium. 2010.

Игры ⚽ Поможем написать курсовую
Synonyms:

Look at other dictionaries:

  • Telescope — Télescope Pour les articles homonymes, voir Télescope (homonymie). Un télescope, (du grec tele signifiant « loin » et skopein signifiant « regarder, voir »), est un instrument optique permettant d augmenter la luminosité ainsi …   Wikipédia en Français

  • Téléscope — Télescope Pour les articles homonymes, voir Télescope (homonymie). Un télescope, (du grec tele signifiant « loin » et skopein signifiant « regarder, voir »), est un instrument optique permettant d augmenter la luminosité ainsi …   Wikipédia en Français

  • Telescope — Tel e*scope, n. [Gr. ? viewing afar, farseeing; ? far, far off + ? a watcher, akin to ? to view: cf. F. t[ e]lescope. See {Telegraph}, and { scope}.] An optical instrument used in viewing distant objects, as the heavenly bodies. [1913 Webster]… …   The Collaborative International Dictionary of English

  • télescope — [ telɛskɔp ] n. m. • 1614; it. telescopio ou lat. mod. telescopium (1611), formé sur le gr. 1 ♦ Instrument d optique destiné à l observation des objets éloignés, et spécialt des astres. ⇒ lunette (astronomique). Lentilles, miroirs de télescope. 2 …   Encyclopédie Universelle

  • Télescope de 3,6 m — Télescope de 3,6 mètres Télescope 3,6 m de l ESO Le télescope Leonard Euler est à droite au premier plan et le télescope 3,6 m de l ESO est au fond …   Wikipédia en Français

  • Telescope — Tel e*scope (t[e^]l [ e]*sk[=o]p), a. Capable of being extended or compacted, like a telescope, by the sliding of joints or parts one within the other; telescopic; as, a telescope bag; telescope table, etc.; now more commonly replaced by the term …   The Collaborative International Dictionary of English

  • Telescope — Tel e*scope, v. i. [imp. & p. p. {Telescoped}; p. pr. & vb. n. {Telescoping}.] To slide or pass one within another, after the manner of the sections of a small telescope or spyglass; to come into collision, as railway cars, in such a manner that… …   The Collaborative International Dictionary of English

  • Telescope — Tel e*scope, v. t. 1. To cause to come into collision, so as to telescope. [Recent] [1913 Webster] 2. to shorten or abridge significantly; as, to telescope a whole semester s lectures into one week. [PJC] …   The Collaborative International Dictionary of English

  • TeleScope — TéléScope Pour les articles homonymes, voir Télescope (homonymie). TéléScope est une télévision associative qui diffusa à Rouen des courts métrages et des reportages sur la culture et la vie associative de la ville. Elle a été diffusé dans trois… …   Wikipédia en Français

  • telescope — [tel′ə skōp΄] n. [It telescopio (coined by GALILEO, 1611) < ModL telescopium < Gr tēleskopos, seeing from a distance: see TELE & SCOPE] an optical instrument for making distant objects, as the stars, appear nearer and consequently larger:… …   English World dictionary

  • telescope — index abridge (shorten), abstract (summarize), constrict (compress) Burton s Legal Thesaurus. William C. Burton. 2006 …   Law dictionary

Share the article and excerpts

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