Tychos.info | Reference Link

Techniques of Astronomy
Astronomy

James Evans for Encyclopedia Britannica
2017



The techniques of astronomy

Astronomical observations involve a sequence of stages, each of which may impose constraints on the type of information attainable. Radiant energy is collected with telescopes and brought to a focus on a detector, which is calibrated so that its sensitivity and spectral response are known. Accurate pointing and timing are required to permit the correlation of observations made with different instrument systems working in different wavelength intervals and located at places far apart. The radiation must be spectrally analyzed so that the processes responsible for radiation emission can be identified.

Telescopic observations

Before Galileo Galilei’s use of telescopes for astronomy in 1609, all observations were made by naked eye, with corresponding limits on the faintness and degree of detail that could be seen. Since that time, telescopes have become central to astronomy. Having apertures much larger than the pupil of the human eye, telescopes permit the study of faint and distant objects. In addition, sufficient radiant energy can be collected in short time intervals to permit rapid fluctuations in intensity to be detected. Further, with more energy collected, a spectrum can be greatly dispersed and examined in much greater detail.

Optical telescopes are either refractors or reflectors that use lenses or mirrors, respectively, for their main light-collecting elements (objectives). Refractors are effectively limited to apertures of about 100 cm (approximately 40 inches) or less because of problems inherent in the use of large glass lenses. These distort under their own weight and can be supported only around the perimeter; an appreciable amount of light is lost due to absorption in the glass. Large-aperture refractors are very long and require large and expensive domes. The largest modern telescopes are all reflectors, the very largest composed of many segmented components and having overall diameters of about 10 metres (33 feet). Reflectors are not subject to the chromatic problems of refractors, can be better supported mechanically, and can be housed in smaller domes because they are more compact than the long-tube refractors.

The angular resolving power (or resolution) of a telescope is the smallest angle between close objects that can be seen clearly to be separate. Resolution is limited by the wave nature of light. For a telescope having an objective lens or mirror with diameter D and operating at wavelength λ, the angular resolution (in radians) can be approximately described by the ratio λ/D. Optical telescopes can have very high intrinsic resolving powers; in practice, however, these are not attained for telescopes located on Earth’s surface, because atmospheric effects limit the practical resolution to about one arc second. Sophisticated computing programs can allow much-improved resolution, and the performance of telescopes on Earth can be improved through the use of adaptive optics, in which the surface of the mirror is adjusted rapidly to compensate for atmospheric turbulence that would otherwise distort the image. In addition, image data from several telescopes focused on the same object can be merged optically and through computer processing to produce images having angular resolutions much greater than that from any single component.

The atmosphere does not transmit radiation of all wavelengths equally well. This restricts astronomy on Earth’s surface to the near ultraviolet, visible, and radio regions of the electromagnetic spectrum and to some relatively narrow “windows” in the nearer infrared. Longer infrared wavelengths are strongly absorbed by atmospheric water vapour and carbon dioxide. Atmospheric effects can be reduced by careful site selection and by carrying out observations at high altitudes. Most major optical observatories are located on high mountains, well away from cities and their reflected lights. Infrared telescopes have been located atop Mauna Kea in Hawaii, in the Atacama Desert in Chile, and in the Canary Islands, where atmospheric humidity is very low. Airborne telescopes designed mainly for infrared observations—such as on the Stratospheric Observatory for Infrared Astronomy (SOFIA), a jet aircraft fitted with astronomical instruments—operate at an altitude of about 12 km (40,000 feet) with flight durations limited to a few hours. Telescopes for infrared, X-ray, and gamma-ray observations have been carried to altitudes of more than 30 km (100,000 feet) by balloons. Higher altitudes can be attained during short-duration rocket flights for ultraviolet observations. Telescopes for all wavelengths from infrared to gamma rays have been carried by robotic spacecraft observatories such as the Hubble Space Telescope and the Wilkinson Microwave Anisotropy Probe, while cosmic rays have been studied from space by the Advanced Composition Explorer.

Angular resolution better than one milliarcsecond has been achieved at radio wavelengths by the use of several radio telescopes in an array. In such an arrangement, the effective aperture then becomes the greatest distance between component telescopes. For example, in the Very Large Array (VLA), operated near Socorro, New Mexico, by the National Radio Astronomy Observatory, 27 movable radio dishes are set out along tracks that extend for nearly 21 km. In another technique, called very long baseline interferometry (VLBI), simultaneous observations are made with radio telescopes thousands of kilometres apart; this technique requires very precise timing.

Earth is a moving platform for astronomical observations. It is important that the specification of precise celestial coordinates be made in ways that correct for telescope location, the position of Earth in its orbit around the Sun, and the epoch of observation, since Earth’s axis of rotation moves slowly over the years. Time measurements are now based on atomic clocks rather than on Earth’s rotation, and telescopes can be driven continuously to compensate for the planet’s rotation, so as to permit tracking of a given astronomical object.

Use of radiation detectors

Although the human eye remains an important astronomical tool, detectors capable of greater sensitivity and more rapid response are needed to observe at visible wavelengths and, especially, to extend observations beyond that region of the electromagnetic spectrum. Photography was an essential tool from the late 19th century until the 1980s, when it was supplanted by charge-coupled devices (CCDs). However, photography still provides a useful archival record. A photograph of a particular celestial object may include the images of many other objects that were not of interest when the picture was taken but that become the focus of study years later. When quasars were discovered in 1963, for example, photographic plates exposed before 1900 and held in the Harvard College Observatory were examined to trace possible changes in position or intensity of the radio object newly identified as quasar 3C 273. Also, major photographic surveys, such as those of the National Geographic Society and the Palomar Observatory, can provide a historical base for long-term studies.

Photographic film converted only a few percent of the incident photons into images, whereas CCDs have efficiencies of nearly 100 percent. CCDs can be used for a wide range of wavelengths, from the X-ray into the near-infrared. Gamma rays are detectable through their Compton scattering, electron-positron pair production, or Cerenkov radiation. For infrared wavelengths longer than a few microns, semiconductor detectors that operate at very low (cryogenic) temperatures are used. Reception of radio waves is based on the production of a small voltage in an antenna rather than on photon counting.

Spectroscopy involves measuring the intensity of the radiation as a function of wavelength or frequency. In some detectors, such as those for X-rays and gamma rays, the energy of each photon can be measured directly. For low-resolution spectroscopy, broadband filters suffice to select wavelength intervals. Greater resolution can be obtained with prisms, gratings, and interferometers. (For additional information on astronomical radiation detectors, see telescope: Advances in auxiliary instrumentation.)

Multi-messenger astronomy

Most of what is known about the universe comes from observations of electromagnetic radiation. However, there are other “cosmic messengers.” Gravity waves are disturbances in space-time that can be detected by very large laser interferometers. Gravity waves and gamma-ray bursts have been observed from neutron-star mergers. Neutrinos and cosmic rays are other particles that can, in principle, be observed; however, as yet, these latter messengers cannot be identified with specific sources. Using two or more of these methods is called multi-messenger astronomy.

Solid cosmic samples

As a departure from the traditional astronomical approach of remote observing, certain more recent lines of research involve the analysis of actual samples under laboratory conditions. These include studies of meteorites, rock samples returned from the Moon, cometary and asteroid dust samples returned by space probes, and interplanetary dust particles collected by aircraft in the stratosphere or by spacecraft. In all such cases, a wide range of highly sensitive laboratory techniques can be adapted for the often microscopic samples. Chemical analysis can be supplemented with mass spectrometry, allowing isotopic composition to be determined. Radioactivity and the impacts of cosmic-ray particles can produce minute quantities of gas, which then remain trapped in crystals within the samples. Carefully controlled heating of the crystals (or of dust grains containing the crystals) under laboratory conditions releases this gas, which then is analyzed in a mass spectrometer. X-ray spectrometers, electron microscopes, and microprobes are employed to determine crystal structure and composition, from which temperature and pressure conditions at the time of formation can be inferred.

Theoretical approaches

Theory is just as important as observation in astronomy. It is required for the interpretation of observational data; for the construction of models of celestial objects and physical processes, their properties, and their changes over time; and for guiding further observations. Theoretical astrophysics is based on laws of physics that have been validated with great precision through controlled experiments. Application of these laws to specific astrophysical problems, however, may yield equations too complex for direct solution. Two general approaches are then available. In the traditional method, a simplified description of the problem is formulated, incorporating only the major physical components, to provide equations that can be either solved directly or used to create a numerical model that can be evaluated (see numerical analysis). Successively more-complex models can then be investigated. Alternatively, a computer program can be devised that will explore the problem numerically in all its complexity. Computational science has taken its place as a major division alongside theory and experiment. The test of any theory is its ability to incorporate the known facts and to make predictions that can be compared with additional observations.

Impact of astronomy

No area of science is totally self-contained. Discoveries in one area find applications in others, often unpredictably. Various notable examples of this involve astronomical studies. Isaac Newton’s laws of motion and gravity (see also celestial mechanics: Newton’s laws of motion) emerged from the analysis of planetary and lunar orbits. Observations during the 1919 solar eclipse provided dramatic confirmation of Albert Einstein’s general theory of relativity, which gained further support with the discovery of the binary pulsar designated PSR 1913+16 and the observation of gravity waves from merging black holes and neutron stars. (See relativity: Experimental evidence for general relativity.) The behaviour of nuclear matter and of some elementary particles is now better understood as a result of measurements of neutron stars and the cosmological helium abundance, respectively. Study of the theory of synchrotron radiation was greatly stimulated by the detection of polarized visible radiation emitted by high-energy electrons in the supernova remnant known as the Crab Nebula. Dedicated particle accelerators are now being used to produce synchrotron radiation to probe the structure of solid materials and make detailed X-ray images of tiny samples, including biological structures (see spectroscopy: Synchrotron sources).

Astronomical knowledge also has had a broad impact beyond science. The earliest calendars were based on astronomical observations of the cycles of repeated solar and lunar positions. Also, for centuries, familiarity with the positions and apparent motions of the stars through the seasons enabled sea voyagers to navigate with moderate accuracy. Perhaps the single greatest effect that astronomical studies have had on our modern society has been in molding its perceptions and opinions. Our conceptions of the cosmos and our place in it, our perceptions of space and time, and the development of the systematic pursuit of knowledge known as the scientific method have been profoundly influenced by astronomical observations. In addition, the power of science to provide the basis for accurate predictions of such phenomena as eclipses and the positions of the planets and later, so dramatically, of comets has shaped an attitude toward science that remains an important social force today. Michael Wulf Friedlander The Editors of Encyclopædia Britannica

History of astronomy

Astronomy was the first natural science to reach a high level of sophistication and predictive ability, which it achieved already in the second half of the 1st millennium bce. The early quantitative success of astronomy, compared with other natural sciences such as physics, chemistry, biology, and meteorology (which were also cultivated in antiquity but which did not reach the same level of accomplishment), stems from several causes. First, the subject matter of early astronomy had the advantage of stability and simplicity—the Sun, the Moon, the planets, and the stars, moving in complex patterns, to be sure, but with great underlying regularity. Biology is far more complicated. Second, the subject was easily mathematized, and already in Greek antiquity astronomy was frequently regarded as a branch of mathematics. This may seem a paradox to a modern reader, since mathematized sciences are regarded as difficult. But in ancient Babylonia and Greece, it was precisely because the motions of the planets could be subjected to mathematical treatment that astronomy made such rapid headway. By contrast, physics failed to make great gains until the 17th century, when its subject matter finally was successfully mathematized. And third, astronomy benefited from its close connection with religion and philosophy, which provided a social value that other sciences simply could not match.

The astronomical tradition is of impressive duration and continuity. A few Babylonian observations of Venus are preserved from the early 2nd millennium bce, and the Babylonians brought their science to a high level by the 4th century bce. For the next half millennium, the greatest headway was made by Greek astronomers, who put their own stamp on the subject but who built on what the Babylonians had accomplished. In the early Middle Ages the leading language of astronomical learning was Arabic, as Greek had been before. Astronomers in Islamic lands mastered what the Greeks had accomplished and soon added to it. With the revival of learning in Europe, and the European Renaissance, the leading language of astronomy became Latin. The European astronomers drew first on Greek astronomy, as translated from Arabic, before acquiring direct access to the classics of Greek science. Thus, modern astronomy is part of a continuous tradition, now almost 4,000 years long, that cuts across multiple cultures and languages. This article focuses on this central story line.

In doing so, there is regrettably little space for other fascinating branches of the history of astronomy. New World astronomy, for example, developed in complete independence but did not rise to so advanced a level. In China astronomy developed to a much higher level, but there too (despite intermittent contacts with Islamic and Indian astronomy and even a fascinating hint of Babylonian influence in the Chinese reckoning of days in 60-day intervals) the story is largely a separate one. That changed with the 16th- and 17th-century Jesuit missions to China, which brought European and Chinese astronomy into direct contact. In India too astronomy reached a high level, involving original Indian methods as well as Indian adaptations of Babylonian and Greek methods, often obtained through Persian contacts. All these branches of the history of astronomy are fascinating and fully merit their own account, but they do not form a part of the main story line of this article.

Prehistory and antiquity

Prehistory

In the French Maritime Alps, in the Vallée des Merveilles (about 100 km [60 miles] north of Nice), are thousands of petroglyphs dating from the Bronze Age (c. 2900–1800 bce). The culture left images of the objects that concerned it—horned animals, the weapons used to hunt them, and so on. There is one clear image of the Sun—a circle with rays coming from it—and, more controversially, archaeologists have identified two images of the star group known as the Pleiades, represented here perhaps by clusters of small cupules carved into the rock. The sky disk of Nebra, a circular bronze plate with areas of applied gold foil, is much clearer as astronomical imagery. It was found in Saxony-Anhalt, Germany, and dates from about 1600 bce. Its golden images include the crescent Moon, probably the Sun (or perhaps the full Moon), and a cluster of seven small gold dots that almost certainly do represent the Pleiades.

Astronomical connections are apparent in a number of prehistoric monuments and graves. In several Stone Age cultures, burial chambers often faced east. Stonehenge (c. 3000–1520 bce) was aligned so that its principal axis coincided with the direction of sunrise on summer solstice. Some other astronomical alignments in Stonehenge, such as with the Moon’s most southerly rising and most northerly setting point, are accepted by many archaeoastronomers. However, most discount some of the more extravagant claims—e.g., that Stonehenge functioned as an eclipse predictor.

That prehistoric people should have noticed and kept track of the Sun and the Moon is not astonishing, but because they lived before writing, the meanings that they attached to celestial events are bound to remain obscure. Some early work in archaeoastronomy was harmed by too great a reliance on conjecture, but methods have greatly improved. Modern archaeoastronomers realize that, with enough stones to work with, one can always find some alignment that is correlated with something celestial. Therefore, one must be careful to perform adequate statistical tests to make sure the alignments are significant and not just accidental.

Mesopotamia

The earliest sophisticated astronomy arose in ancient Babylonia, in central Mesopotamia, and there are three reasons why it happened there rather than, say, in ancient Greece. First, in Babylonia astronomy had an important social function: the gods sent signs from heaven to warn the king about impending war, a bad barley harvest, or an impending epidemic. In the early 2nd millennium bce, the pattern of taking celestial omens was already established. This was long before the rise of personal astrology; whereas common people might have taken signs from their surroundings—for example, by observing the behaviour of animals—the celestial signs were intended for the king and kingdom alone. The Greeks were no less superstitious than any other ancient people and saw omens in the flight of birds, in dreams, or in the frenzied utterances of an oracle, but they had no early custom of celestial divination. That came later, in the Hellenistic period, after contact with Babylonian wisdom. Second, there was in Babylonia a civil service charged with things astronomical. Temple scribes, who were often priests, watched the sky every night to keep track of what transpired, and they recorded their observations. Third, in Mesopotamia there existed a stable technology for recording data—the clay tablet. As long as they are protected from water, clay tablets are practically indestructible. The acquired data also had a secure place for storage (the temples), and broken tablets were recopied. All of these circumstances—a social function, a bureaucracy charged with doing astronomy, and a secure system for data storage—were missing in the early Greek world.

By the 7th century bce, astronomical diaries were in existence. These recorded the results of night-by-night watching by the temple astronomers, such as when a planet passed by the Pleiades or another reference star, when Venus reemerged from its period of invisibility (after having been too near the Sun), or when Jupiter stood still and went into retrograde motion (that is, reversed direction). These ancient Babylonian observations were not very precise, but it is far more important to have a long run of observations than to have precise ones.

Within a few generations, Babylonian astronomers had achieved the ability to predict the behaviour of the Moon and the planets. Though no planet repeats its motion from one year to the next, repetition does occur if one waits long enough. For example, Venus does not go into retrograde in the same month or in the same sign of the zodiac from one retrogradation to the next. The pattern does not repeat until after 5 complete retrograde cycles, which take about 8 years. Similarly, Mars starts a new repeating pattern of retrogradations after 22 cycles (which take 47 years), and Saturn repeats its pattern after 57 retrogradations (59 years). This discovery gave rise to the Babylonian goal-year texts. Supposing that one wanted to predict the behaviour of all the planets for the year 2025, which would be the goal year, one could look back in the records and find what Venus had done in 2017 (8 years earlier), what Mars had done in 1978 (47 years earlier), and so on. Thus, the first predictive planetary astronomy was achieved with a good database by making use of repeating patterns.

By about 300 bce the temple scribes achieved a far more sophisticated method of predicting planetary behaviour on the basis of complex arithmetical theories. For each planet there are several different versions of the planetary theory preserved. The basic idea was that a key event, such as the onset of retrogradation, could be thought of as an object in its own right that worked its way around the zodiac. For example, in one version of the Babylonian theory, Jupiter’s onsets of retrograde motion were spaced at regular intervals of 30° though about half the zodiac (in Jupiter’s slow zone) but at 36° intervals in the remainder of the zodiac (Jupiter’s fast zone). A scribe could use this theory to rapidly work out the dates and positions in the zodiac of the onsets of Jupiter’s retrograde motion for a century or more.

Ancient Greece

Astronomy is present from the beginning of Greek literature. In Homer’s Iliad and Odyssey, stars and constellations are mentioned, including Orion, the Great Bear (Ursa Major), Boötes, Sirius, and the Pleiades. More-detailed astronomical knowledge is found in Hesiod’s Works and Days, from perhaps a generation later than Homer. Hesiod used the appearances and disappearances of important fixed stars in the course of the annual cycle in order to prescribe the work to be done around the farm or the seasons for safe sailing. Much of the astronomical knowledge in Hesiod paralleled the knowledge of the contemporary Babylonians, but the Greeks were substantially less advanced.

Applying geometry

The breakthrough that gave Greek astronomy its own particular character was the application of geometry to cosmic problems. The oldest extant source that clearly states that Earth is a sphere and that gives a sound argument to support the claim is Aristotle’s On the Heavens (c. 350 bce), but this knowledge likely went back several generations earlier. Aristotle mentioned that Earth’s shadow as seen on the Moon during a lunar eclipse is circular. He also mentioned the changes that occur in the stars that are visible as one moves from north to south on Earth. Aristotle stated that certain mathematicians had contrived to measure Earth’s circumference and had found a value of 400,000 stades. Although stades of several different lengths were in use, a typical stade was about 0.18 km (0.11 mile), which means that a value for Earth’s circumference was about 72,000 km (44,000 miles). (The true value is 40,075 km [24,902 miles].) Although it is not known who made the first such measurement, Aristotle may have been referring to Eudoxus of Cnidus, whom Aristotle knew in Athens and who wrote a book (now lost) called The Circuit of the Earth.

The famous measurement by Eratosthenes (the oldest measurement of the size of Earth for which details survive) was made in the 3rd century bce. Eratosthenes used the fact that at noon on the summer solstice, the Sun was directly overhead in Syene (a town on the upper Nile, at modern Aswan, Egypt), but in Alexandria on the same day, the Sun was below the vertical by about one-fiftieth of a circle (7.2°). This, together with an estimate of 5,000 stades for the distance between Alexandria and Syene, gave a value of 50 × 5,000 = 250,000 stades (about 45,000 km, or 28,000 miles) for the circumference of Earth, a figure that was roughly correct, regardless of the exact value of Eratosthenes’ stade.

Also in the 3rd century bce, Aristarchus of Samos applied geometrical reasoning to estimate the distances of the Sun and the Moon, in On the Sizes and Distances of the Sun and Moon. However, his initial premises included several questionable numerical values. For example, he assumed that at the moment of quarter Moon, the angle between the Sun and the Moon, as observed from Earth, is 87°. From this it followed that the Sun’s distance is about 19 times the Moon’s distance from us. (The actual ratio is about 389.) A second doubtful observation was that the angular size of the Sun or the Moon is 2° (the actual value is about 0.5°). Although the numerical inputs were flawed, Aristarchus’s method was valid. He found the Moon’s diameter to be between 0.32 and 0.4 times the diameter of Earth, and the Sun’s diameter to be between 6.3 and 7.2 times the diameter of Earth. (The diameters of the Moon and the Sun compared with that of Earth are actually 0.27 and 109, respectively.) By the time of Hipparchus of Bithynia (2nd century bce), improvements on Aristarchus’s method had led to excellent values for the size and distance of the Moon. But the ancients always considerably underestimated the size and distance of the Sun, which is so far from Earth that a measurement of its parallax lay beyond the powers of naked-eye astronomy. Aristarchus’s 19-to-1 ratio was not called seriously into question until the 17th century.

The motion of the planets

Greek thinking about the motion of the planets began by about 400 bce. Eudoxus of Cnidus constructed the first Greek theory of planetary motion of which any details are known. In a book, On Speeds (which is lost but was briefly discussed by Aristotle and Simplicius), Eudoxus regarded each celestial body as carried on a set of concentric spheres, which nest one inside another. For each planet, three different motions must be accounted for, and Eudoxus proposed to do this with four spheres. The daily revolution to the west is accounted for by the outermost sphere (1). Next inside is sphere 2, whose axle fits into sphere 1 at an offset of about 24°; sphere 2 turns to the east in the planet’s zodiacal period (12 years for Jupiter, 30 years for Saturn). The third motion is retrograde motion. For this, Eudoxus used a combination of two spheres (3 and 4). The planet itself rides on the equator circle of sphere 4. The axle of 4 fits inside sphere 3 with a slight angular offset. Spheres 3 and 4 turn in opposite directions but at the same speed. The motion of the planet resulting from the gyrations of spheres 3 and 4 is a figure eight, which lies in the spherical surface. Eudoxus likely understood the mathematical characteristics of this curve, as he gave it the name hippopede (horse fetter). The two-sphere assembly of 3 and 4 is inserted into the inner surface of sphere 2. Thus, all three motions are accounted for, at least qualitatively: the daily motion to the west by sphere 1, the slow motion eastward around the zodiac by sphere 2, and the occasional retrograde motion by the two-sphere assembly of 3 and 4. Eudoxus’s theory is sometimes called the theory of homocentric spheres, as all the spheres have the same centre, Earth.

At this stage, Greek astronomers were more interested in providing plausible physical accounts of the universe and in proving geometrical theorems than in providing numerically accurate descriptions of planetary motion. Eudoxus’s successor Callippus made some improvements to the model. Nevertheless, the homocentric spheres were criticized for their failure to account for the fact that some planets (notably Mars and Venus) are much brighter at some times of their cycles than at others. Eudoxus’s system was soon abandoned as a theory for the motion of the planets, but it exerted a profound influence in cosmology, for the cosmos continued to be regarded as a set of concentric spheres until the Renaissance.

Late in the 3rd century bce, alternative theoretical models were developed, based on eccentric circles and epicycles. (An eccentric circle is a circle that is slightly off-centre from Earth, and an epicycle is a circle that is carried and rides around on another circle.) This innovation is usually attributed to Apollonius of Perga (c. 220 bce), but it is not conclusively known who first proposed these models. In considering the Sun’s motion, Eudoxus’s theory of homocentric spheres ignored the fact that the Sun appears to speed up and slow down in the course of the year as it moves around the zodiac. (This is clear from spring’s being several days longer than fall.) An eccentric (i.e., off-centre) circle can explain this fact. The Sun is still considered to travel at constant speed around a perfect circle, but the centre of the circle is slightly displaced from Earth. When the Sun is closest to Earth, it appears to travel a little more rapidly in the zodiac. When it is farthest away, it appears to travel a little more slowly. As far as is known, Hipparchus was the first to deduce the amount and direction of the off-centredness, basing his calculations on the measured length of the seasons. According to Hipparchus, the off-centredness of the Sun’s circle is about 4 percent of its radius. The eccentric-circle theory was capable of excellent accuracy in accounting for the observed motion of the Sun and remained standard until the 17th century.

The standard theory of the planets involved an eccentric circle, which carried an epicycle. Imagine looking down on the plane of the solar system from above its north pole. The planet moves counterclockwise on its epicycle. Meanwhile, the centre of the epicycle moves counterclockwise around the eccentric circle, which is centred near (but not quite exactly at) Earth. As viewed from Earth, the planet will appear to move backward (that is, go into retrograde motion) when it is at the inner part of the epicycle (closest to Earth), for this is when the westward motion of the planet on the epicycle is more than enough to overcome the eastward motion of the epicycle’s centre forward around the eccentric.

Hipparchus played a major role in introducing Babylonian numerical parameters into Greek astronomy. Indeed, an important shift in Greek attitudes toward astronomy occurred about this time. The Babylonian example served as a sort of wake-up call to the Greeks. Previous Greek planetary thinking had been more about getting the right big picture, based on philosophical principles and geometrical models (whether using Eudoxus’s concentric spheres or Apollonius’s epicycles and eccentrics). The Babylonians had no geometrical models but instead focused on devising arithmetical theories that had real predictive power. Hipparchus achieved numerically successful geometrical theories for the Sun and the Moon, but he did not succeed with the planets. He contented himself with showing that the planetary theories then in circulation did not agree with the phenomena. Nevertheless, Hipparchus’s insistence that a geometrical theory, if it is true, ought to work in detail marked a major step in Greek astronomy.

Another of Hipparchus’s contributions was the discovery of precession, the slow eastward motion of the stars around the zodiac caused by wobbling, over a period of 25,772 years, in the orientation of Earth’s axis of rotation. Hipparchus’s writings on this subject have not survived, but his ideas can be reconstructed from summaries given by Ptolemy. Hipparchus used observations of several fixed stars, taken with respect to the eclipsed Moon, which had been made by some of his predecessors. On comparing these with eclipse observations he had made himself, he deduced that the fixed stars move eastward not less than 1° in 100 years. The Babylonians, in their theories, revised their locations of the equinoxes and solstices. For example, in one version of the Babylonian theory, the spring equinox is said to occur at the 10th degree of Aries; in another version, at the 8th degree. Some historians have maintained that this reflects a Babylonian awareness of precession, on which Hipparchus might have drawn. Other historians have argued that the evidence is not clear and that these differing norms for the equinox may represent nothing more than alternative conventions.

Ptolemy

The culminating work of Greek astronomy is the Almagest of Claudius Ptolemaeus (2nd century ce). Ptolemy built on the work of his predecessors—notably Hipparchus—but his work was so successful that it made older works of planetary astronomy superfluous, and they ceased to be read and copied. An innovation that appears for the first time in the Almagest is the equant point. As in the planetary theories of Hipparchus’s day, a planet travels uniformly around its epicycle while the centre of the epicycle moves around Earth on an off-centre circle. But in Ptolemy’s theory the motion of the epicycle’s centre is nonuniform—it speeds up and slows down—which was a radical departure from Aristotelian physics. However, the nonuniformity is expressed in the language of uniformity: the epicycle’s centre moves in such a way that it appears to go through equal angles in equal times as viewed from another point distinct from Earth, the equant point. Though this may seem like an unnecessary complication, it was just what an explanation of planetary motion required (for, in the modern view, planets really do move nonuniformly). In Ptolemy, for the first time, Greek geometrical planetary theory finally achieved real numerical accuracy. Ptolemy’s theory actually predicted the behaviour of the planets, and it dominated the practice of astronomy for 1,400 years.

The Almagest contains an account of the observations and a description of the mathematical procedures that Ptolemy used to deduce the parameters of his theories. It also provides tables that allow the user to work out the position of a planet from theory for any desired date. The advantage of the tables is that Ptolemy has done all the trigonometry. One need only follow Ptolemy’s precepts, take numbers out of the various tables, and combine them to get an answer for a planet’s position. The Almagest includes trigonometric tables and a catalog of about 1,000 stars, which was probably based substantially on an earlier catalog by Hipparchus but with additions and modifications by Ptolemy. It also contains Ptolemy’s improvement on Hipparchus’s lunar theory. As an aid to convenient calculation, Ptolemy also composed the Procheiroi kanones (Handy Tables), in which the astronomical tables of the Almagest were expanded and accompanied by directions for using them but were stripped of the theoretical discussion.

Historians have long debated how much credit to give Ptolemy and how much to assign to his predecessors. For an ancient scientist, he was unusually generous in crediting his predecessors, particularly Hipparchus, for discoveries. But he did not always mention the origin of his ideas. In any case, Ptolemy’s publications fundamentally changed the way astronomy was done in the Greek world. In the period between Hipparchus and Ptolemy, Greek astronomers had struggled without great success to make geometrical planetary theory work. From the evidence of Greek-inspired astronomical works that later turned up in India, it has been conjectured that Greek astronomers before Ptolemy may have experimented with nonuniform motion—something akin to the equant—but nothing remains of a finished project before Ptolemy.

The garbage dumps of Oxyrhynchus in Greek Egypt have yielded large quantities of papyri, including planetary tables used for computing horoscopes. Most of this material is from the 1st through the 4th century ce. The papyri show the astrologers of Greek Egypt happily using Greek versions of Babylonian arithmetical theories for computing planet positions. This material is found side by side with papyri based on Ptolemy’s Handy Tables. Thus, in Greek astronomy there was a high road based on philosophy of nature and rooted in geometrical methods, and there was a low road based on the convenient arithmetical methods adapted from the Babylonians, even if these could not be considered to rest on adequate physical or philosophical foundations. These two methods still existed side by side up to the time of Ptolemy, and even a little after, but the newly successful (and convenient) geometrical methods gradually won out.

Ptolemy also wrote a speculative cosmological work, the Hypotheseis ton planomenon (Planetary Hypotheses), in which he took the eccentric-and-epicycle astronomy of the Almagest as physically true. However, to give a satisfactory image of the cosmos, he needed the nested-spheres cosmology of Eudoxus. The eccentrics and epicycles were regarded as the equator circles of three-dimensional orbs. Assuming that there was no wasted or empty space in the cosmos (consistent with both Aristotelian and Stoic physics), Ptolemy supposed that the mechanism for Mercury must lie immediately above the mechanism for the Moon. The mechanism for Venus came next, and so on, out to the mechanism for Saturn, and finally the sphere of the fixed stars. (Historians are not all agreed on whether the ancients regarded these mechanisms as real, physical objects that moved the planets or as merely theoretical constructions, but Ptolemy probably considered them as real things.) The known distance of the Moon provided the scale. When the numbers were worked out, the distance of the fixed stars was about 20,000 Earth radii. This is an enormous cosmos (though much smaller than modern estimates). Accepting such a conclusion, based on planetary astronomy and a few auxiliary physical premises, required a certain courage of imagination.

India, the Islamic world, medieval Europe, and China

India

Ptolemy was the last major figure in the Greek astronomical tradition. Commentaries were written on his works by Pappus of Alexandria in the 3rd century ce and by Theon of Alexandria and his daughter, Hypatia, in the 4th, but creative work was no longer being done. Babylonian astronomy traveled eastward into Persia and India, where it was adapted in original ways and combined with native Indian methods. Greek geometrical planetary theories, from the time between Hipparchus and Ptolemy, also made their way into India. This material is of great complexity and variety and is difficult to sort out. For example, Babylonian arithmetical procedures used for computing lunar and solar phenomena turn up in conjunction with a length for the solar year due to Hipparchus. Nevertheless, the Indian material, besides its own intrinsic interest, provides information about Greek astronomy during a vital period about which the Classical texts say little.

The Islamic world

In the 8th century, Arabic Muslim astronomers came into contact with this complicated astronomical material. Theories and methods that had passed from the Babylonians and Greeks through Persia to India now came back to the West. A good example is provided by the zīj of Muḥammad ibn Mūsā al-Khwārizmī (9th century). Al-Khwārizmī’s work is a confusing mixture of Indian, Persian, and Greek tables and techniques, but it helped establish an important genre of the zīj. A zīj is a handbook of astronomical tables, including tables for working out positions of the Sun, Moon, and planets, accompanied by directions for using them. The ancient prototype was Ptolemy’s Handy Tables.

Ptolemy’s Almagest was translated on at least four occasions into Arabic. Much of the translation activity centred on the Baghdad caliphate of the ʿAbbāsids (750–1258). With the pure geometrical form of Greek planetary theory now available, Arabic astronomers worked to master it and then to improve upon it. The zīj of al-Battānī (early 10th century ce) showed mastery of Ptolemaic planetary theory and improved values for some of Ptolemy’s parameters, such as the magnitude and direction of the Sun’s eccentricity. Hundreds of Arabic zījes from the 9th to the 15th century have been preserved. Some were based on Indian methods, but the great majority were in the tradition of the Almagest and the Handy Tables. A zīj that was very influential in the development of European astronomy was the Toledan Tables, compiled in Spain by a group of Muslim and Jewish astronomers, put into final form by Ibn al-Zarqallu around 1080, and translated into Latin soon after. (The Toledan Tables are mentioned by Chaucer in The Canterbury Tales.)

With the passage of time, it became possible for astronomers to make new discoveries, including those that depended on detecting slow changes in the heavens. In the 9th century the Baghdad astronomers observed that the obliquity of the ecliptic had decreased from the value given in Ptolemy’s Almagest. The obliquity of the ecliptic is the angle between the celestial equator and the tropic of Cancer. It corresponds to the northward displacement of the Sun between the equinox and the summer solstice and can be measured by means of noon altitudes of the Sun taken at key times of the year. Between Ptolemy’s time and the present day, the obliquity of the ecliptic has decreased by about a quarter of a degree. Arabic astronomers also noted that the seasons had changed slightly in length from the values recorded by Ptolemy. This implied that the solar apogee has a slow motion to the east. Thus, the centre of the Sun’s circle can be regarded as revolving very slowly about Earth. This motion was represented in al-Battānī’s zīj.

Ptolemy’s planetary theory was criticized, but minor disagreements between Ptolemy’s tables and actual observations of the planets did not play a significant role in this criticism. Most of the criticism centred on Ptolemy’s violation of the Aristotelian principle of the uniformity of the celestial motions. About 1000 ce Ibn al-Haytham criticized the equant point in Shukūk ʿalā Baṭlamyūs (“Doubts About Ptolemy”). Ibn al-Haytham also objected to Ptolemy’s habit of defining motions with respect to immaterial points and lines as if they were real material bodies. (Complaints about the artificiality of Ptolemy’s constructions had been made even in late antiquity—for example, by the Greek philosopher Proclus in his Diadochi hypotyposis astronomicarum positionum [“Sketch of Astronomical Hypotheses”].)

Ibn al-Haytham’s doubts about Ptolemaic planetary theory inspired some creative mathematical modeling by 13th-century astronomers associated with the observatory of Marāgheh (now in Azerbaijan). Naṣīr al-Dīn al-Ṭūsī described a construction through which two circular motions can give rise to the oscillation of a point back and forth along a straight line. Ptolemy’s theories of Mercury and of the Moon involved oscillatory movements for which the standard mechanisms seemed philosophically questionable. Al-Ṭūsī applied his two-circle mechanism (called an “al-Ṭūsī couple” by modern scholars) to produce the same phenomena in what seemed to him a physically more plausible way. Al-Ṭūsī’s student al-Shīrāzī went farther, using a minor epicycle to eliminate the need for an equant point. In the 14th century Ibn al-Shāṭir of Damascus built on the works of the Marāgheh school in his Nihāyat al-suʾl fi taṣḥīḥ al-uṣūl (“Final Inquiry Concerning the Rectification of Planetary Theory”), which was also characterized by the elimination of nonuniform motions in favour of minor epicycles. However, these efforts did not transform common practice, since the overwhelming majority of late medieval planetary tables are Ptolemaic in their underlying theory. In the 16th century Nicolaus Copernicus used models identical to those of Ibn al-Shāṭir and the Marāgheh school. How he came by them is unknown, but there are too many of them to make independent discovery credible. These technical “improvements” on Ptolemy had nothing to do with the heliocentric hypothesis, but they show that Copernicus was heir to a tradition of critical engagement with Ptolemy.

Medieval Europe

In the Latin West the level of scientific learning had sunk to a low level. None of the Greek works most important for ancient astronomy and cosmology—Aristotle’s On the Heavens and Ptolemy’s Almagest, Handy Tables, and Planetary Hypotheses—were available. The teaching of astronomy was based on a number of low-level Latin accounts. Book II of Pliny the Elder’s Naturalis historia (Natural History, 1st century ce) contained a summary of astronomical matters. In the 4th century Martianus Capella wrote an allegorical poem, De nuptiis Philologiae et Mercurii (The Marriage of Philology and Mercury). In the two introductory books, Philology, personified as a maiden, is wed to Mercury, patron god of learning. In the following seven books, each of the Liberal Arts, including Astronomy, personified as a handmaid to Philology, steps forward to give an account of her art. Martianus’s Marriage was widely admired in the early Middle Ages as a compendium of all useful learning.

In the 12th and 13th centuries, two developments were key to the revival of astronomy in the Latin West. The translation movement rapidly made available key works of Greek astronomy that had long been out of reach. One of the most important translators was Gerard of Cremona. As his students later wrote of him, he had learned everything that was known to the Latins, but for love of the Almagest, which he had heard of but which was unavailable in Latin, he went to Spain and learned Arabic well enough to translate it. Thus, one could maintain that a major reason for the revival of learning in the West was one man’s desire to be able to read Ptolemy. Gerard translated from Arabic versions not only Ptolemy’s Almagest but also Aristotle’s On the Heavens, Euclid’s Elements, and about two dozen other works of astronomy and geometry. In a single generation most of the key works of ancient astronomy became available.

The second important development was the foundation of the European universities, starting with those of Bologna, Paris, and Oxford. Because astronomy figured among the liberal arts, it had a place in the university core curriculum. Of course, the astronomy of the liberal arts curriculum was at a rudimentary level. The students might be taken though an introduction to the celestial sphere—for example, the De sphaera mundi (“On the Sphere of the World,” c. 1230) by Johannes de Sacrobosco, which might be followed by the anonymous Theorica planetarum (“Theories of the Planets”), a superficial introduction to eccentrics and epicycles. Nevertheless, in every university town there had to be someone charged with teaching astronomy.

In the 1270s a new set of astronomical tables was compiled in Spain under the patronage of the Christian king Alfonso X of Léon and Castile. These were based on standard Ptolemaic astronomy, with some differences in the treatment of precession (now considered to occur at a variable speed). By 1320 the Alfonsine Tables had reached Paris, where they were reworked by several Parisian astronomers. From there they spread all over Latin Europe, and for more than two centuries they were the standard.

China

Though “oracle bones” exist from the late 2nd millennium bce that mention observations of lunar and solar eclipses as well as the appearance of a new star (nova), astronomical reports begin to be fairly numerous only from about 200 bce. In China astronomy had an imperial function. The emperor was considered the Son of Heaven. Thus, the regulation of the calendar, as well as the success or failure of his astronomers to predict an eclipse, reflected either well or badly on him. Many different astronomical summaries were written in conjunction with the ascent of a new emperor. Usually these emphasized the lunisolar calendar, but later they also included tables for predicting the motions of the planets, as well as eclipses. Chinese predictive astronomy used repeating arithmetical cycles and was thus more like Babylonian astronomy than like Greek astronomy. Perhaps because the Chinese were less tied up with cosmological theories and “laws” of nature than the Greeks and their medieval European successors were, the Chinese astronomers were much more interested in singular events, such as comets, novae, meteor showers, solar eclipses, and sunspots (which the Chinese discovered before the Europeans), and they kept detailed records of them.

Renaissance

European astronomy regained the level of the ancient Greeks only with the publication in 1496 of Epytoma in Almagestum Ptolemaei (“Epitome of Ptolemy’s Almagest”) begun by mathematician and astronomer Georg von Peuerbach and completed by his student Regiomontanus (the Latin name of Johannes Müller von Königsberg). Regiomontanus’s chapter-by-chapter commentary helped the next couple of generations learn their Ptolemy. He sometimes criticized Ptolemy—for example, pointing out that the twofold variation in the distance of the Moon implied by Ptolemy’s lunar theory greatly exceeded the variation in distance implied by the Moon’s variation in apparent size. Although this variation had been known in Arabic astronomy, this was its first mention in the Latin West.

Copernicus

Polish astronomer Nicolaus Copernicus announced the motion of Earth in De revolutionibus orbium coelestium libri VI (“Six Books Concerning the Revolutions of the Heavenly Orbs,” 1543). (An early sketch of his heliocentric theory, the Commentariolus, had circulated in manuscript in the small astronomical community of central Europe from about 1510, but it was not printed until the 19th century.) Although Copernicus made some new observations of the planets and drew on some observations by his medieval predecessors, new observations played no important role in his discovery. Rather, Copernicus discovered the motion of Earth by understanding Ptolemy more deeply than anyone else had—for the essential clues lay there in the Almagest for all to see.

Each planet’s motion is connected with the motion of the Sun. The inferior planets are always the close companions of the Sun. Mercury never gets more than about 22° from the Sun, and Venus never more than about 48°. This can be explained simply by imagining that these two planets circle the Sun.

For the superior planets (Mars, Jupiter, and Saturn), the connection is more subtle. Each of these planets goes into retrograde motion when it is diametrically opposite the Sun as viewed from Earth. In the ancient planetary theory, this required the three planets to move around their epicycles in lockstep with one another and with the motion of the Sun around Earth. In the case of Mars, for example, the revolving line from the epicycle’s centre to Mars must remain parallel to the revolving line from Earth to the Sun. The same holds true for Jupiter and Saturn. Ptolemy mentioned that one could use this fact to avoid duplicated calculations if one wanted to work out the positions of all three planets for the same date. Copernicus’s great insight was that these four simultaneous motions were really manifestations of one single motion—the motion of Earth itself.

The early reaction to Copernicus was rather muted, and astronomers had several different kinds of response. One could admire Copernicus’s mathematical abilities and simply remain agnostic on the question of Earth’s motion. Such, for example, was the position of German astronomer Erasmus Reinhold, who wrote a popular textbook of Ptolemaic astronomy but who also computed and published the Prutenic Tables, based on Copernicus’s planetary theory, which helped boost Copernicus’s reputation.

Tycho

Danish astronomer Tycho Brahe was a good example of those who admired Copernicus’s achievement in tying all the motions of the planets more closely to the Sun but who were unable to accept the motion of Earth. Brahe worked out an alternative cosmology, known as the Tychonic system. In this view the Moon and the Sun revolve around Earth, but all of the other planets revolve around the moving Sun. Tycho’s system had the same explanatory advantages as Copernicus’s. It was what the Copernican system would look like if Earth was made to stay at rest.

Like many other astronomers, Tycho was fascinated by the brilliant new star that appeared in Cassiopeia in 1572. He made extensive observations to determine if it shifted its position with respect to neighbouring stars from night to night. For an astronomer or a philosopher of an Aristotelian frame of mind, it would be difficult to admit that a new star really could appear in the heavens; one would more likely consider it to be some sort of phenomenon in the upper reaches of the air and fire (the elements that, in Aristotle’s cosmology, surround Earth and the seas). If the new star displayed a parallax (i.e., if it shifted back and forth with respect to the real stars), one could be sure that it was near Earth and not a part of the cosmic sphere. Tycho’s demonstration that the new star had no measurable parallax and that it therefore really was a star in the celestial sphere did much to dismantle the old physics.

In 1577 there was a second gift from heaven—a particularly bright comet. In antiquity and the Middle Ages, comets were regarded as atmospheric phenomena. Thus, Aristotle did not treat them in On the Heavens but rather treated them in Meteorology. After all, they are transient, appear suddenly, rapidly cross from one constellation to another, and then disappear. However, Tycho was able to model the motion of the comet by putting it into an orbit around the Sun. He pointed out that the comet was therefore sometimes closer to Earth, and sometimes farther away, than Venus and Mercury were. This seemed to imply that it crashed through the celestial spheres that carried these planets, thus calling into question these vast constructions.

Galileo

In 1609 Italian scientist Galileo Galilei, using his own telescope, modeled on an invention recently made in the Netherlands, discovered that the Moon, far from being smooth and utterly unlike Earth, had mountains and craters. By using the lengths of their shadows, Galileo was even able to measure the heights of the Moon’s mountains. A number of nebulae resolved into swarms of individual small stars. Even the Milky Way was made of stars. Perhaps the most exciting find was the discovery of four moons revolving about Jupiter. These discoveries were announced in Galileo’s Sidereus Nuncius (The Sidereal Messenger, 1610), the book that made his reputation. Although none of these discoveries directly supported the Copernican theory, they all lent indirect support in that they made the new cosmology less objectionable. That Jupiter has satellites cannot prove that Earth goes around the Sun, but it showed that there was at least one other centre of revolution than Earth. It also showed that a moving planet could carry its satellites along with it (as Earth does the Moon in the Copernican view). The later discovery that Venus ran through a complete set of phases like the Moon definitely ruled out the Ptolemaic idea that Venus lay below the solar sphere, but it did not rule out a theory like Tycho Brahe’s, in which Venus circled the Sun while the Sun moved around Earth.

In 1616 the Roman Catholic Church placed Copernicus’s De revolutionibus on the Index Librorum Prohibitorum (Index of Forbidden Books). In 1618 a list was issued of 10 specific corrections—passages dealing with Earth’s motion that were to be struck out. But outside Italy, this order was rarely followed. One curious exception can be mentioned. Jesuit missionaries to China in the late 16th and early 17th centuries carried European astronomy with them. The Jesuit astronomers were predisposed to the Tychonic system, which kept Earth at the centre of the universe but which otherwise shared the advantages of Copernicanism. After the condemnation of Copernicanism, they had no choice but to keep to the Tychonic system, and they continued to teach it in China long after it had gone out of fashion in Europe.

Galileo, even though he had been warned not to teach Copernicanism as literally true, decided to take advantage of the ascent of a more liberal thinker to the papacy, Pope Urban VII, and gambled on a work of popularization. In Dialogo sopra i due massimi sistemi del mondo, tolemaico e copernicano (Dialogue Concerning the Two Chief World Systems, Ptolemaic & Copernican, 1632), three friends—an avowed follower of Ptolemy and Aristotle, a convinced Copernican, and an intermediary who guided the debate—discuss cosmology and astronomy. Naturally, the Copernican gets the better of the arguments. Galileo was put on trial and forced to recant, but the official condemnation of the book and the trial of Galileo did little to halt the advance of the new ideas.

Kepler

German astronomer Johannes Kepler embraced Copernicanism wholeheartedly. Nevertheless, he may be considered the last astronomer, and one of the greatest astronomers, in the old tradition—one of the last for whom the Almagest was still a part of the research literature. Kepler had begun with an interest in cosmology, in trying to understand God’s architecture for the solar system. Why, for example, were there six planets rather than some other number? Going back to speculations introduced by Plato and the Pythagoreans, Kepler applied first the geometry of the regular solids and then musical harmonies to explain various aspects of the universe. For example, there were six planets because there were only five regular solids (cube, tetrahedron, etc.), which God had used as spacers between the planetary orbs when working out the cosmic architecture. Kepler’s first book, Mysterium cosmographicum (“Cosmographic Mystery,” 1596), was based on this idea. As a result of this book, Kepler received an invitation to work with Tycho Brahe, but nothing happened until 1600, when Tycho left his native Denmark and relocated to Prague under the patronage of the Holy Roman emperor Rudolf II.

Kepler went to Prague, hoping to obtain from Tycho better values of planetary parameters so that he could refine his cosmology. The collaboration lasted only a short time, because Tycho died in 1601. When Kepler arrived in Prague, Tycho and his assistants were involved in observations of Mars, which was then about to make a near approach to Earth. This turned out to be fortunate for Kepler, because only Mars and Mercury have large enough eccentricities to make the departures of their orbits from circularity appreciable and Mercury is too near the Sun to be easily or often observed. After Tycho’s death Kepler gained access to his observation records. Far from being able to find ready results to use in cosmology, Kepler was forced to analyze many observations to put them into usable form.

Kepler began as a convinced Copernican, so he put the Sun in the middle of his system, but for technical details he went back to Ptolemy. Kepler began by regarding Mars as moving on a circle that was slightly off-centre from the Sun and was following Ptolemy’s equant law. But he was unable to get this theory to match all of Tycho’s observations to better than about 8 minutes of arc (1 minute of arc = 1/60 of a degree), and he believed that Tycho’s observations were good to about 2 minutes of arc. Against his will he was forced to reexamine the fundamentals of planetary motion. This led to the first two of Kepler’s laws of planetary motion, published in Astronomia Nova (New Astronomy, 1609). According to the first law, the paths of planets are ellipses with one focus located at the Sun. The second law, which was actually discovered first, makes a small improvement on Ptolemy’s equant: a planet moves around the Sun at a variable speed in such a way that the line from the Sun to the planet sweeps out equal areas in equal times. In Harmonice Mundi (The Harmony of the World, 1619), Kepler announced his third, or harmonic, law: the ratio a3/T2 is the same for all planets, where a is the semimajor axis of a planet’s elliptical orbit and T is the orbital period.

Enlightenment

Newton

Kepler’s laws received a physical explanation only with the publication of English physicist and mathematician Isaac Newton’s Philosophiae Naturalis Principia Mathematica (Mathematical Principles of Natural Philosophy, 1687). Here Newton announced his laws of motion, as well as the law of universal gravitation: any two particles in the universe attract one another with a force proportional to the product of their masses and inversely proportional to the square of the distance between them. Newton used these laws to rederive Kepler’s laws, thus making planetary theory a branch of physics for the first time in history. He then applied the laws to explain other phenomena, such as the rise and fall of the tides and the orbits of comets.

The law of inertia (Newton’s first law—a body tends to move at constant speed in a straight line) had been hinted at by Galileo and expressed in a more definite way by French philosopher René Descartes. The third law (if body A exerts a force on body B, then B exerts force on A equal in magnitude but opposite in direction) was well supported by recent work on collisions by Dutch mathematician Christiaan Huygens and others. Newton’s second law (the force impressed on a body is equal to the body’s mass times its acceleration) represented a fresh way of thinking about motion. The idea of an inverse-square law for gravity had been toyed with in England by physicist Robert Hooke, architect Sir Christopher Wren, and astronomer Edmond Halley, but they had been unable to assemble all the necessary concepts—the law of attraction, the concept of motion under an impressed force, and the linking mathematics—into a finished product. Newton’s Principia fundamentally altered the intellectual context for the science of astronomy.

Newton’s law of universal gravitation encountered some resistance, especially on the French-speaking Continent, where it was sometimes regarded as a falling back into a discredited way of thinking. The idea that one body could reach out across empty space and affect another seemed to some to be a throwback to medieval animism. It did not help that Newton could not explain the mechanism by which gravity acted.

Testing Newton’s theory

In the first part of the 18th century, the inverse-square law was subjected to several dramatic tests. The first concerned Earth’s shape. Newton had argued that Earth’s rapid rotation on its axis must cause Earth to depart from perfect sphericity. Instead, Earth should be an oblate spheroid—that is, flattened at the poles like an onion. For evidence, Newton pointed to the example of Jupiter, which showed a noticeable flattening when seen through a telescope. Also, in 1672 the French scientist Jean Richer had carefully measured the rate of a pendulum clock near Earth’s Equator (by comparing it with the motion of the stars) and found that the clock ran slightly slower than an identical clock in Paris. Newton argued that if Earth was flattened at the poles, Paris would be a little closer to Earth’s centre than the Equator would be. If gravity varied as the inverse square of the distance, Earth’s gravity should then be stronger in Paris than at the Equator, and thus the Paris pendulum clock would run faster. But in 1718 Jacques Cassini announced results of a survey of the Paris meridian from Dunkirk to Collioure, made by his father, Gian Domenico Cassini, and himself, that seemed to show just the opposite—that Earth is elongated at the poles like a lemon. French natural philosophers, steeped in the vortex theory of Descartes, found ways of explaining this in terms of Cartesian physics. In the 1730s the French Academy of Sciences sponsored two expeditions—one to Lapland, led by mathematician Pierre-Louis Moreau de Maupertuis, and one to equatorial South America—just to settle this question. Careful geodetic and astronomical measurements were made to determine the length of a degree of the meridian for a place near the pole and a place near the Equator. The results of the Lapland expedition showed decisively that Earth was flattened at the poles, as Newton had maintained. Voltaire famously addressed his friend Maupertuis as “the flattener of the world and of Cassinis.”

Second, Newton had been unable to calculate the correct rate for the advance of the Moon’s perigee—that is, the movement of the point on the Moon’s orbit where it is closest to Earth. The reason for the advance of the perigee lies in the perturbing attraction of the Sun on the Moon, but Newton obtained a rate too small by half (a complete revolution of the perigee takes about 18 years instead of the observed 9). In the 18th century several leading mathematicians tried to solve the problem and failed. In 1747 French mathematician and physicist Alexis-Claude Clairaut proposed a modification of Newton’s law of gravity. Instead of a pure inverse-square law, Clairaut proposed adding a small term, proportional to the inverse fourth power of the distance, in order to get the motion of the Moon’s perigee to come out correctly. Clairaut later withdrew this proposal and showed in a new calculation that the inverse-square law was perfectly adequate for explaining the motion of the Moon’s perigee. The problem was too complex to be solved directly, and it was necessary to introduce approximations. Clairaut showed that the approximations made by Newton and those who followed had been too rash and that with more-careful approximations, the advance of perigee came out just right. This was, by far, the most precise test of the Newtonian theory to date.

Finally, as the time approached for the expected reappearance of Halley’s Comet, celestial mechanicians undertook a more-precise calculation of the date of return. Halley had argued that the comets of 1531, 1607, and 1682 were one and the same and predicted a return for late 1758 or early 1759, but he did not live to see it happen. When the comet, on its very elongated orbit, passes by massive planets, such as Jupiter, on its way out of and back into the inner solar system, the planets exert forces that perturb its motion. In Paris, Clairaut, astronomer Jérôme Lalande, and Nicole Lepauté, the wife of a well-known instrument maker, calculated the motion of the comet, including the perturbing forces. This was the most ambitious program of numerical integration ever undertaken up to that time. When the comet reappeared within their announced one-month window of error, it was seen by many as a triumph of calculation, as well as of the law of universal gravitation.

Laplace

Since every planet is attracted not only by the Sun but also (much more weakly) by all the other planets, its orbit cannot really be the simple ellipse described by Kepler. Newton was therefore willing to entertain the idea that God might occasionally need to readjust the planetary system. In the 18th century new mathematical methods were developed, largely in France, to treat perturbations more efficiently. The key figures in this work were Joseph-Louis Lagrange and Pierre-Simon Laplace. They showed that the solar system is inherently quite stable. Each planet is perturbed by the others, but the net result is only oscillatory corrections to the unperturbed orbits; there are no runaway behaviours. God would not need to intervene after all.

Laplace is known mainly for his densely mathematical Traité de mécanique céleste (A Treatise of Celestial Mechanics; 5 vol., 1798–1825), but he was also the author of a work of popularization, the Exposition du système du monde (The System of the World), which appeared in several editions between 1796 and 1824. In this work Laplace explained for the lay reader all the phenomena of the solar system in terms of universal gravitation. This was followed by a brief history of astronomy from ancient times down to Laplace’s own day. The book ended with a brief account of what is now called Laplace’s nebular hypothesis, a theory of the origin of the solar system. Laplace imagined that the planets had condensed from the primitive solar atmosphere, which originally extended far beyond the limits of the present-day system. As this cloud gradually contracted under the effects of gravity, it first formed rings and then amalgamated into planets. Newton had seen in the regularities of the solar system a sure sign of the wisdom and beneficence of the Creator. For example, the fact that all the planets travel around the Sun in the same direction and more or less in the same plane could be explained only by divine providence. Laplace, looking at the same facts, instead regarded them as evidence about the prehistory of the solar system. The nebular hypothesis, although only sketchily worked out, was important as an early example of an evolutionary theory in natural science, and it is notable that evolutionary thinking entered astronomy before it became important in the life sciences.

The age of observation

Herschel and the new planet

The foremost observational astronomer of the period was William Herschel. Herschel was born in Hannover, Germany, in 1738, but he moved to England as a young man to avoid the Continental wars. He settled in Bath and made a living as a musician and as a music teacher while devoting all his spare time to amateur astronomy, which he cultivated at a very high level. By making his own telescopes, he soon had finer instruments than anyone else. In 1781, while sweeping the sky for double stars, he spotted a small object that he first took to be either a comet or a nebulous star. Herschel convinced himself that he had discovered a new comet, which would not have been an unusual occurrence, but soon other astronomers demonstrated that it was moving in a nearly circular orbit about the Sun, so it became known as a planet. Angling for royal patronage, Herschel proposed naming the new object Georgium Sidus, the Georgian Star, after King George III. The flattery worked, for Herschel was soon rewarded with an annual pension, which allowed him to give up teaching music and to devote himself almost completely to astronomy. Continental astronomers refused to accept Herschel’s proposed name. In 1783 German astronomer Johann Elert Bode proposed Uranus, the name that eventually stuck.

There was a long tradition that went all the way back to Plato and the Pythagoreans of trying to tie planetary distances to numerical sequences. An influential new scheme was proposed in 1766 by Prussian astronomer Johann Daniel Titius von Wittenberg. According to Titius, the sequence of planetary distances takes the form 4, 4 + 3, 4 + 6, 4 + 12, 4 + 24, 4 + 48, 4 + 96,… Titius fixed the scale by assigning 100 to Saturn’s distance from the Sun which, indeed, makes Mercury’s distance about 4. Titius pointed out that there is an empty place at distance 28, corresponding to the large gap between Mars and Jupiter, and speculated that this gap would be filled by undiscovered satellites of Mars. Titius had slipped his distance rule, unsigned, into his German translation of Swiss philosopher Charles Bonnet’s Contemplation de la nature (The Contemplation of Nature, 1764). This sequence of planetary distances was adopted, without credit, by Bode in his Deutliche Anleitung zur Kenntniss des gestirnten Himmels (“Clear Guide to the Starry Heaven”; 2nd ed., 1772). (In later editions Bode did give credit to Titius.) Bode also predicted that a planet would eventually be found at distance 28. Herschel’s discovery of Uranus at distance 192 (where the Titius-Bode sequence predicted 196) seemed an uncanny confirmation of the law.

Astronomers began to search for a planet in the Mars-Jupiter gap. In 1801 Italian astronomer Giuseppe Piazzi discovered a small planetlike object in the gap, which he named Ceres, after the patron goddess of Sicily. Pallas was discovered by German astronomer Wilhelm Olbers the following year. Herschel did not feel that these objects were large enough to be planets, so he proposed the term asteroid (Greek for “starlike”), which had been suggested to him by classicist Charles Burney, Jr., via his father, music historian Charles Burney, Sr., who was a close friend of Herschel’s. (Later they were also called “minor planets.” Today, after a 2006 ruling by the International Astronomical Union, they are officially designated “dwarf planets” if, like Ceres, they are massive enough to have been rounded to spheres by their own gravity. Most, however, are much smaller and are officially designated “small solar system bodies,” though many astronomers still informally refer to these as asteroids.)

Herschel and the Milky Way

Although Herschel’s discovery of Uranus made his reputation, it was far from being his most important contribution. During the 18th century, astronomers had measured the proper motions of a reasonably large number of stars. (Proper motion is the slow drift of a star with respect to its neighbours, which slowly causes the constellations to change shape. The first few proper motions were announced in 1718 by Halley, who found them by comparing recently observed star positions with data recorded in Ptolemy’s Almagest.) Herschel noted that many of the stars with substantial proper motions are bright, which suggests that they might be nearby. He reasoned that if there is any pattern in the stellar proper motions, it might be due to the motion of the Sun through the field of stars. In 1783 Herschel published an analysis of 19 proper motions and concluded that the Sun is traveling through space in the direction of the constellation Hercules (toward a point called the solar apex). This was later questioned by the 19th-century German astronomer and mathematician Friedrich Bessel, who had many more proper motions to work with, but Herschel’s conclusion was ultimately proved correct.

Herschel was not only an excellent observer but also a remarkably inventive thinker in devising simplifying assumptions that allowed him to make theoretical progress. He was one of several people who in the mid- to late 18th century arrived at the idea that the Milky Way has the form of a flattened disk. However, only Herschel actually tried to deduce the structure of this vast star system. If one had a telescope sufficiently powerful to penetrate to the edge of the Milky Way and aimed the telescope in a direction lying in the plane of the Milky Way, one would look through a region dense with stars. In fact, the number of stars seen in the telescope’s field of view could be taken as a measure of the distance from the Sun to the edge of the Milky Way in that direction. Herschel made a large number of such counts, which he called “star gages,” and in 1785 drew the first quantitative chart of the Milky Way’s form. Later, when he realized that his telescope had not actually been powerful enough to penetrate to the galaxy’s edge, he abandoned this drawing. But because there was nothing that might replace it, Herschel’s drawing of the form of the Milky Way was frequently reprinted throughout the 19th century.

Herschel and his sister Caroline Herschel expended prodigious time and effort in cataloging the nebulae. A few small nebulous, or cloudlike, patches in the night sky are visible to the naked eye and had been mentioned by ancient Greek and medieval Arabic astronomers. In 1755 German philosopher Immanuel Kant suggested that these nebulae might be vast systems of stars, comparable to the Milky Way. These later came to be called “island universes,” but at this stage the notion was purely speculative. Nebulae could be troublesome, since astronomers on the lookout for new comets could easily mistake an uncataloged nebula for a comet. In 1771 French astronomer Charles Messier published a list of 45 nebulae to keep himself and other comet searchers from wasting time. In 1784 his list was expanded to 103. These Messier objects are today favourite objects for amateur astronomers. William Herschel received a copy of one of Messier’s lists. Caroline, who swept for comets by using a special telescope that William had made for her, soon noticed nebulae not on Messier’s list. As a consequence, William became interested in nebulae and systematically searched for them while engaged in other observing chores. Over 20 years he raised the number of known nebulae to about 2,500. It had been known since Galileo that through a good telescope, some nebulae could be resolved into stars. Were the nebulae really all star systems at vast distances from Earth, or were there also regions of true nebulosity, clouds of luminous fluid? Recent drawings of the Orion nebula, when compared with a drawing made by Christiaan Huygens in the 17th century, seemed to show that this nebula had changed form, which implied that it had to be close, relatively small, and not made of stars. Herschel’s opinions changed in the course of his career, but he tended to regard nebulae as star systems in the process of evolution toward denser states, with the evolution driven by universal gravitation. Today it is known that “nebulae” come in several kinds: some are clouds of glowing gas; some are clusters of stars; and some really are galaxies comparable in size to the Milky Way. But this understanding was not possible until the 19th-century development of spectroscopy and the 20th-century measurement of the distance to another galaxy.

Herschel was unusual among the astronomers of his day, because he concerned himself with the larger construction of the heavens and was far less interested in the ordinary business of professional astronomy, which meant making exactingly accurate position measurements. Herschel helped to open the road to a new physical astronomy that really came into its own only in the 20th century. Nevertheless, there were important discoveries to be made with the old style of astronomy, conducted at universities or sponsored by national observatories, ostensibly because of its application to navigation.

English astronomer James Bradley was perhaps the most significant of the old-style 18th-century observers. High-precision measurements of star positions that he made in the 1720s led him to the discovery of the aberration of starlight. It had been known since the late 17th century that light has a finite speed. Danish astronomer Ole Rømer used that idea in 1676 to explain why the eclipses of the satellites of Jupiter appear to run alternately about 10 minutes ahead of or behind schedule over the course of a year. Christiaan Huygens then worked out a numerical estimate for the speed of light that was published in his Traité de la Lumière (Treatise on Light, 1690). Bradley discovered that the fixed stars too suffer apparent annual changes in their positions. For example, a star near the pole of the ecliptic (seen at right angles to the plane of Earth’s orbit) appears to execute a small circular motion, of 20 seconds of arc radius, in the course of a year. A second discovery by Bradley introduced an even more troublesome complication—the nutation (or nodding) of Earth’s axis, which has an amplitude of about 8 seconds. Because of aberration and nutation, the fullest possible precision of astronomical observations could not be achieved unless the observations were corrected for these effects.

Precise calculations and observations

A major aspect of 19th-century astronomy was the move toward greater precision both in methods of calculation and in quantitative methods of observation. Here the natural successor to Bradley was Friedrich Wilhelm Bessel, who reduced Bradley’s enormous collection of star positions for aberration and nutation and in 1818 published the results in a new star catalog of unprecedented accuracy, the Fundamenta Astronomiae (“Foundations of Astronomy”).

No better demonstration of improved methods could be wished for than the near-simultaneous measurements of stellar parallaxes by Friedrich Georg Wilhelm von Struve of the star Vega in 1837, by Bessel of the star 61 Cygni in 1838, and by Scottish astronomer Thomas Henderson of the triple star Alpha Centauri in 1838. The annual parallax is the tiny back-and-forth shift in the direction of a relatively nearby star, with respect to more-distant background stars, caused by the fact that Earth changes its vantage point over the course of a year. Since the acceptance of Copernicus’s moving Earth, astronomers had known that stellar parallax must exist. But the effect is so small (because the diameter of Earth’s orbit is tiny compared with the distance of even the nearest stars) that it had resisted all efforts at detection. For example, the parallax of 61 Cygni is 0.287 seconds of arc (1 second of arc = 1/3,600 of a degree). The shift from parallax was observed only after the development of precise astronomical instruments, such as the heliometer that German physicist and optician Joseph von Fraunhofer built for Bessel, that could measure stellar positions to the necessary accuracy of hundredths of a second of arc. (In the preceding century Bradley, who could measure stellar positions only with an accuracy of half a second of arc, had been making a failed attempt to detect stellar parallax when he stumbled instead on the aberration of light.) The successful measurement of stellar parallaxes gave for the first time accurate values for the distances of stars other than the Sun.

By about 1820 it was clear that Uranus was not keeping to the schedule of motion predicted for it. In the 1840s John Couch Adams in England and Urbain-Jean-Joseph Le Verrier in France independently sought to explain the anomaly through the gravitational attraction of an undiscovered planet outside the orbit of Uranus. Both Adams and Le Verrier assumed the rough validity of the Titius-Bode law to make their calculations easier. Adams predicted a place in the zodiac where astronomers should look, but at first he could not get the English astronomical community to tackle the job. Le Verrier had better luck, for his prediction was taken up immediately by Johann Gottfried Galle at the Berlin Observatory, who found the new planet Neptune in 1846, near the place in the sky where Le Verrier said it would be. This episode caused a stormy period in English-French scientific relations, as well as recriminations in the English astronomical community for the failure to pursue Adams’s prediction in a timely way.

In Ireland a wealthy amateur, William Parsons, 3rd earl of Rosse, inspired by Herschel’s example, continued the quest for larger and better telescopes. Because Herschel had treated the optics of his large telescopes as trade secrets, Rosse had to do all his own design by trial and error. In 1839 Rosse built a 36-inch (91-cm) reflecting telescope, with the mirror made of polished metal, and then, in 1845, the 72-inch (183-cm) “Leviathan of Parsonstown.” That year, using this gigantic instrument, Rosse observed and sketched the spiral form of the nebula known as Messier 51. Three years later he sketched the spiral shape of Messier 99. Rosse and his helpers eventually described more than 60 spiral nebulae.

The rise of astrophysics

In 1835 the French positivist philosopher Auguste Comte cited the chemical constitution of the stars as an example of knowledge that might be forever hidden. However, unknown to Comte, the development of spectroscopy was already revealing the composition of the stars and permitting the emergence of a true astrophysics. In 1802 English physician William Hyde Wollaston saw several dark gaps or lines in the Sun’s spectrum and conjectured that these might be the natural boundaries between colours. The dark lines in the solar spectrum were rediscovered around 1814 in Munich by Fraunhofer, who cataloged some 500 of them. Fraunhofer noted that his dark D line in the yellow part of the solar spectrum matched up with the well-known bright line in the spectrum of a candle flame. Fraunhofer also showed that light from Venus shows the same structure as sunlight, and he observed dark lines in the spectra of a number of bright stars.

A key step was taken in 1849 by French physicist Jean Foucault, who showed that the bright orange lines seen in the light emitted by a carbon arc could also be observed as dark absorption lines in sunlight that was passed through the gas around the arc. Thus, a gas that can be stimulated to emit a particular colour will also preferentially absorb that same colour. Around 1859 German chemist Robert Wilhelm Bunsen and physicist Gustav Robert Kirchhoff showed how to associate spectral lines with particular chemical elements. From an analysis of the dark lines in the solar spectrum, Kirchhoff concluded that iron, calcium, magnesium, sodium, nickel, and chromium were present in the Sun. In 1868 English astronomer Joseph Norman Lockyer identified an orange line in a solar-prominence spectrum that had no counterpart in that of any known element, so he ascribed it to a new element, which he called helium (after helios, the Greek name for the Sun and the Sun god). Helium was not isolated on Earth until 1895 by Scottish chemist William Ramsay.

In the 1860s Italian astrophysicist Angelo Secchi described the spectra of some 4,000 stars and classified them into four groups. A star’s spectrum is continuous, with all the colours present, though it may be brighter in one or another part of the spectrum according to the temperature of the star. (Cooler stars are redder.) Typically, the continuous spectrum is also overlaid with a number of dark absorption lines. Secchi’s classification scheme was based on the overall colour of the star, the number and kind of absorption lines, and other features of the spectrum. This work, performed before the application of photography to spectroscopy, was slow and very tedious.

Also in the 1860s English astronomer William Huggins observed the spectrum of a bright nebula and found that it consisted only of bright emission lines. This was therefore a glowing gas—a case of true nebulosity. Huggins went on to observe about 70 nebulae. He found that the nebulae consisted of two major groups. About one-third were gaseous, and about two-thirds showed the continuous spectrum that would be expected of unresolved stars.

A major centre of spectroscopy in the next generation was the Harvard College Observatory, under the direction of American astronomer Edward Charles Pickering. By putting a prism in front of the object lens of a telescope, his team was able to photograph the spectra of many stars at once. The resulting Henry Draper Catalogue (named to recognize the financial support for the project provided by Draper’s widow) appeared in nine volumes between 1918 and 1924 and contained over 225,000 spectra. Key to this work was a new stellar-classification scheme (still in use today—for example, the Sun is a G-type star) refined by American astronomer Annie Jump Cannon, who had joined Pickering’s team in 1895.

In the mid-19th century there was considerable dispute about the reality and nature of the Doppler effect. A shift in the frequency of light received from a moving source had been proposed in 1842 by the Austrian physicist Christian Doppler, who (wrongly) thought that in this way he could explain the colours of binary stars. The Doppler effect was demonstrated for sound by the Dutch physicist Christophorus Henricus Didericus Buys-Ballot in 1845 by putting musicians on a moving train. In 1868 Huggins measured a small shift in the position of the F line in the hydrogen spectrum for Sirius, which was interpreted as being caused by the radial motion of the star with respect to Earth. Strong confirmation of the Doppler effect for light was obtained in the 1870s by German astronomer Hermann Karl Vogel, who measured the spectral shift between the east and west edges of the rotating Sun. In the 1880s Vogel and German astronomer Julius Scheiner began to measure the radial velocities of stars by using photographic spectra. The tabulation of spectral types and radial velocities soon became a standard part of star cataloging.

The cataloging of stellar spectra opened the way for new discoveries, for it soon became clear that the spectral type of a star has a relation to the star’s intrinsic brightness. However, since a star will look dimmer the farther away it is, the intrinsic brightness (or absolute magnitude) of a star cannot be known unless one first has a way to determine the distance. American astronomer Henry Norris Russell in 1913 published a scatter plot correlating absolute magnitude with spectral type, using only stars for which he judged that the distances had been well determined. Slightly earlier, German astronomer Hans Rosenberg and Danish astronomer Ejnar Hertzsprung had plotted similar diagrams, using only stars from a single cluster, either the Pleiades or the Hyades. (Stars in a single cluster are all at roughly the same distance from Earth, so their apparent magnitudes can be used as replacements for their absolute magnitudes.) The resulting scatter plots are called Hertzsprung-Russell (H-R) diagrams. The H-R diagram revealed that most stars lie on a “main sequence,” in which absolute magnitude is positively correlated with temperature. Bluer main-sequence stars (spectral type O or B) are much brighter than main-sequence red stars (spectral type K or M). The H-R diagram also showed a second branch, in which there are reddish stars that are much brighter than those on the main sequence. If these bright red stars have the same surface temperature (because they are of the same spectral type) as a main-sequence star but are much brighter, they must be physically larger, and they soon came to be called “red giants.” White dwarfs were soon discovered as yet another branch. The H-R diagram became crucial for guiding speculations about the evolution of stars.

The source of the energy that drives the stars had been a great mystery. In the 19th century, chemical combustion and heating due to gravitational contraction were the only possibilities, but Scottish physicist William Thomson (Lord Kelvin) pointed out that a chemical process could hardly last more than 3,000 years. In various versions of heating by release of gravitational energy, the Sun was supposed to be contracting slowly (by about 75 metres [246 feet] per year) or else be heated by the continual infall of meteoric matter. After the discovery of radioactivity in the 1890s and the realization that Earth’s interior was warmed by this mechanism, various schemes were proposed for explaining stellar energy in terms of radioactive decay. The true explanation came only after German American physicist Albert Einstein’s 1905 publication of the mass-energy relation (E = mc2, a consequence of special relativity). In the 1920s English astrophysicist Arthur Eddington proposed the proton-proton reaction, in which four atoms of hydrogen are combined to produce one atom of helium, with the mass difference released in the form of energy. Because of the primitive state of nuclear physics at the time, he could not say in detail how this might occur, but he pointed to the mere existence of helium in the stars as the surest proof that such a process must exist. Nuclear physics gained a firm foundation in the early 1930s with the discovery of the neutron and of deuterium (a heavy isotope of hydrogen with a proton and a neutron in its nucleus). From then on, progress was rapid. In 1937 German physicist Carl Friedrich von Weizsäcker discovered the CNO cycle, in which carbon, nitrogen, and oxygen act as catalysts in a sequence of nuclear reactions that leads to the conversion of hydrogen into helium. In 1939 German American physicist Hans Bethe published a more detailed and quantitative study of the CNO cycle that finally put stellar astrophysics on a secure footing. Bethe also treated in detail the proton-proton reaction that Eddington had only guessed at. In a collision at high temperature, two protons may stay close enough together for the brief time required for one of them to be converted into a neutron by emission of a positron; thus, deuterium is formed. From deuterium, helium may then be built up in several different ways. Bethe also showed that the CNO cycle is more important in high-temperature stars and the proton-proton reaction more important in cooler stars. Nuclear physics was successfully integrated with what was known about the conditions of temperature and density in the interiors of stars.

Relativity

Einstein develops the theory

A key theoretical development for 20th-century astronomy and cosmology was the development of the theory of relativity, from 1905 to 1915, which eventually led to an explanation of the origin of the universe. The theory of relativity grew out of contradictions between electromagnetic theory (worked out by Scottish physicist James Clerk Maxwell in the 1860s) and what people thought they knew about relativity of motion. As a high-school student, Albert Einstein formulated a clever thought experiment and reasoned to a contradiction. According to electromagnetic theory, a light wave consists of changing electric and magnetic fields. Einstein asked what one would see if one could run at the speed of light alongside a light wave. In the frame of reference of the runner, the light wave would be stationary. Hence its fields would not be changing, and consequently the light wave should not exist. Clearly, something was wrong.

There were also experimental difficulties. At the close of the 19th century, it was widely believed that light needs a medium to propagate in (as do other wave disturbances, such as sound). This all-pervading medium was given a name, the luminiferous (or light-carrying) ether. However, several experiments, including the famous Michelson-Morley experiment of 1887, failed to detect any motion of Earth with respect to the ether. Many physicists regarded this as a crisis, and some sought rather desperate ways out. Irish physicist George Francis FitzGerald proposed in 1889 that a moving body, owing to its interaction with the ether, undergoes a contraction in the direction of its motion in just the right amount to explain the null result of the Michelson-Morley experiment. A similar idea was worked out in greater detail by the Dutch physicist Hendrik Antoon Lorentz.

Einstein’s approach was far more fundamental. In his 1905 paper “Zur Elektrodynamik bewegter Körper” (“On the Electrodynamics of Moving Bodies”), he proceeded axiomatically. He assumed, first, that all uniformly moving reference frames are equally valid for doing physics and, second, that the speed of light is always the same, regardless of the relative motion of the source and the receiver. The first postulate was unexceptional; it was a logical result of the scientific revolution from Copernicus to Galileo. The second postulate was unusual (and one can note how it resolves the paradox of the young Einstein: one simply cannot run alongside a light wave so that it appears stationary). From these two postulates, surprising consequences followed: time runs at different rates in different frames of reference; lengths are contracted in the direction of motion (not, as Lorentz and FitzGerald thought, owing to an interaction with the ether but rather owing to the very nature of space and time); and finally, the mass of a moving body increases without limit as the body’s speed approaches that of light. All this is a part of the special theory of relativity (special because gravity is not included and only uniformly moving bodies are considered).

Many of the same equations had been obtained earlier by Lorentz. In France too mathematician Henri Poincaré was tapping at the door. Relativity was in the air in 1905, and if Einstein had not written his paper, the same basic results would have been obtained. Einstein’s merit was his clarity of thought and his axiomatic approach, which showed that something fundamental was at stake. The ether simply had no place in his worldview. Shortly afterward Einstein’s former teacher mathematician Hermann Minkowski formulated the notion of space-time, in which time is regarded as a fourth dimension. This did not change the equations of relativity theory, but it greatly changed the way people thought about the theory and it helped prepare the way for the general theory.

If the special theory of relativity was bound to happen, the same cannot be said of the general theory of relativity, which is a theory of the gravitational field. Einstein turned to the problem of gravitation shortly after publishing his special theory. By 1907 he had already become convinced that a gravitational field should deflect a ray of light. In developing his new theory, he was guided by two considerations. The first was the principle of equivalence—that acceleration and gravitation are somehow manifestations of the same thing. For example, as Einstein put it, a freely falling man would not feel his own weight. The second consideration was general covariance—namely, the assumption that the laws of physics should take the same mathematical forms in arbitrary frames of reference and not just in uniformly moving frames. The mathematics was difficult, and after years of groping, in 1915 Einstein found the general theory of relativity.

Testing relativity

Einstein had little to go on in the way of observational evidence—only the anomalous advance of the perihelion of Mercury. If Mercury were the only planet orbiting the Sun, then, according to Newtonian physics, the orbit would be a perfect ellipse that would always preserve the same orientation. However, the other planets weakly attract Mercury and disturb its orbit so that the long axis of Mercury’s orbital ellipse is not stationary but rotates around the Sun. The point of the orbit where Mercury is closest to the Sun is called the perihelion; thus, the perihelion slowly advances, in the same direction that Mercury moves.

In 1859 Le Verrier announced that the perihelion of Mercury was advancing a little too quickly to be explained by the action of the known planets. The excess was tiny, about 38 seconds of arc per century, compared with the 527 seconds of arc per century that Le Verrier attributed to known planetary perturbations. Le Verrier had discovered Neptune by means of anomalies in the motion of Uranus. Therefore, he naturally guessed that the discrepancy in the motion of Mercury was due to an undiscovered ring of asteroids, or perhaps a planet, lying between Mercury and the Sun. The planet even acquired a name, Vulcan, but soon proved to be illusory. In 1895 American astronomer Simon Newcomb made a fresh study of the problem and confirmed the unexplained anomaly in the motion of the perihelion of Mercury, now in the amount of 43 seconds of arc per century.

Before he published his new theory, Einstein checked that it gave the right answer for the Mercury problem. In the theory of general relativity, Newton’s law of gravity, in which the gravitational force between two bodies decreases with the inverse square of the distance between them, is not completely accurate in describing massive bodies that are very near each other; rather, the law must be modified by a term that decreases with the inverse fourth power of the distance. The mass of the Sun is not very great, and no planets, aside from Mercury, move very near it; thus, in describing the solar system, Newton’s law of gravity had been quite a successful approximation.

A second prediction of Einstein’s gravity theory was a value for the bending of a ray of starlight as it passes by the Sun. Here Einstein had a bit of luck. In a preliminary version of the theory, the bending came out too small by half. Einstein had tried to get astronomers interested in detecting the bending by looking for apparent shifts in the locations of stars near the Sun during a total solar eclipse. However, for various reasons, including the intervention of World War I, no one had succeeded in making a test. Had astronomers been able to test this early prediction, they would have found it erroneous, which could well have impeded acceptance of the final (1915) version of the theory.

In England, Eddington was instrumental in spreading interest in Einstein’s general theory of relativity. Eddington mastered the mathematics, wrote about the subject, and got his colleagues interested in testing the theory. Eddington and Sir Frank Dyson, the astronomer royal, persuaded the Royal Astronomical Society to mount two expeditions to observe the total solar eclipse of 1919, one to Brazil and the other, led by Eddington, to Principe Island, off the west coast of Africa. Stars near the Sun were photographed during totality, when the Sun’s disk was covered by the Moon, and their positions were compared with earlier photographs made of the same part of the sky. The looked-for effect was small, but Eddington and Dyson confirmed that Einstein’s prediction of the bending of light was correct. Their announcement caused an international sensation.

The third effect of general relativity predicted by Einstein was the gravitational redshift. Light coming from a compact massive object should be slightly redshifted; that is, the light should have a longer wavelength. Measuring this was a delicate business, as the expected shift was small and could easily be masked by other effects. Attempts to measure the gravitational redshift by using absorption lines in the solar spectrum led to contradictory and inconclusive results, but in 1925 American astronomer Walter Adams, at Mount Wilson Observatory, announced that he had determined the gravitational redshift of Sirius B, the white dwarf companion of Sirius. (White dwarfs were expected to have much higher gravitational redshifts than stars like the Sun.) The confirmation of the gravitational redshift not only bolstered general relativity but also helped support Eddington’s theory of stellar structure, which predicted enormous densities (and therefore very strong surface gravities) for white dwarfs. It was not realized until decades later that Adams’s measurement of the gravitational redshift of Sirius B was too small by a factor of four. Compensating errors in Adams’s measurements and in Eddington’s ideas of the temperature and radius of Sirius B had produced a fortuitous agreement (Sirius B is even smaller than Eddington thought). This was another case of Einstein’s being lucky, for a conflict between the measured redshift and the theoretical value could again have compromised the acceptance of general relativity. In the case of Adams’s measurements, a major source of trouble was light from Sirius itself scattered into the Sirius B spectrum by Adams’s instruments. However, by then the effect had been successfully measured in laboratories on Earth with just the value that Einstein’s theory predicts.

In Einstein’s theory of gravity, a massive object produces a curvature of the space-time around it. Einstein made his own early predictions of observable phenomena by using partial and approximate solutions. Thus, Einstein was surprised when German astronomer Karl Schwarzschild in 1916 found an exact solution for Einstein’s field equations for the space-time around a spherical body. For the so-called exterior solution, the Schwarzschild space-time presents mild corrections to the Newtonian motion of bodies (such as that which explains the advance of the perihelion of Mercury). But lurking in Schwarzschild’s solution (the so-called interior solution) are signs of a much stranger regime. If a body is so dense that it is confined to a region of radius less than 2GM/c2 (where G is the Newtonian constant of gravitation, M is the mass of the object, and c is the speed of light), the solution becomes problematic: at this critical radius (called the Schwarzschild radius), the solution seems to produce a singularity, where certain mathematical expressions become either zero or infinite. This aspect of Schwarzschild’s solution caused enormous confusion for decades, and astrophysicists tried to paper over the problems by ignoring the interior part of the solution or by seeking coordinate systems in which the problems went away. But in 1939 American physicists J. Robert Oppenheimer and Hartland Snyder published a study of what happens when a star has exhausted its stores of nuclear energy and begins to collapse. Basing their relativistic analysis on the Schwarzschild solution, Oppenheimer and Snyder showed that the Schwarzschild radius does not correspond to a singularity but defines a surface from which light cannot escape to infinity. With collapse, they wrote, the star closes itself off from any communication with an outside observer. This paper helped inaugurate the study of black holes, though these exotic objects did not really come into their own until the 1960s. (The term black hole was coined only in 1967, by American physicist John Archibald Wheeler.) The first plausible candidates for black holes were observed in the 1970s.

Galaxies and the expanding universe

Einstein almost immediately applied his gravity theory to the universe as a whole, publishing his first cosmological paper in 1917. Because he was not well acquainted with recent work in astronomy, he assumed that the universe was static and unchanging. Einstein assumed that matter was distributed uniformly throughout the universe, but he could not find a static solution to his field equations. The problem was that the mutual gravitation of all the matter in the universe would tend to make the universe contract. Therefore, Einstein introduced an additional term containing a factor Λ, the “cosmological constant.” The new term provided a universal cosmic repulsive force, which could act at great distances to counteract the effects of gravity. When he later learned of the expansion of the universe, Einstein described the cosmological constant as the greatest blunder of his career. (But the cosmological constant has crept back into late 20th-century and 21st-century cosmology. Even when Einstein was wrong, he was often onto something profound.)

Einstein’s static solution represented a universe of finite volume but with no edges, as space curved back on itself. Thus, an imaginary traveler could travel forever in a straight line and never come to an edge of the universe. The space has positive curvature, so the angles in a triangle add up to more than 180°, though the excess would be apparent only in triangles of sufficient size. (A good two-dimensional analogy is Earth’s surface. It is finite in area but has no edge.)

At the beginning of the 20th century, most professional astronomers still believed that the Milky Way was essentially the same thing as the visible universe. A minority believed in a theory of island universes—that the spiral nebulae are enormous star systems, comparable to the Milky Way, and are scattered through space with vast empty distances between them. One objection to the island-universe theory was that very few spirals are seen near the plane of the Milky Way, the so-called Zone of Avoidance. Thus, the spirals must somehow be a part of the Milky Way system. But American astronomer Heber Curtis pointed out that some spirals that can be viewed edge-on obviously contain huge amounts of dust in their “equatorial” planes. One might also expect the Milky Way to have large amounts of dust throughout its plane, which would explain why many dim spirals cannot be seen there; visibility is simply obscured at low galactic latitudes. In 1917 Curtis also found three novae on his photographs of spirals; the faintness of these novae implied that the spirals were at great distances from the Milky Way.

The static character of the universe was soon challenged. In 1912, at the Lowell Observatory in Arizona, American astronomer Vesto M. Slipher had begun to measure the radial velocities of spiral nebulae. The first spiral that Slipher examined was the Andromeda Nebula, which turned out to be blueshifted—that is, moving toward the Milky Way—with a velocity of approach of 300 km (200 miles) per second, the greatest velocity ever measured for any celestial object up to that time. By 1917 Slipher had radial velocities for 25 spirals, some as high as 1,000 km (600 miles) per second. Objects moving at such speeds could hardly belong to the Milky Way. Although a few were blueshifted, the overwhelming majority were redshifted, corresponding to motion away from the Milky Way. Astronomers did not, however, immediately conclude that the universe is expanding. Rather, because Slipher’s spirals were not uniformly distributed around the sky, astronomers used the data to try to deduce the velocity of the Sun with respect to the system of spirals. The majority of Slipher’s spirals were on one side of the Milky Way and receding, whereas a few were on the other side and approaching. For Slipher, the Milky Way was itself a spiral, moving with respect to a greater field of spirals.

In 1917 Dutch mathematician Willem de Sitter found another apparently static cosmological solution of the field equations, different from Einstein’s, that showed a correlation between distance and redshift. Although it was not clear that de Sitter’s solution could describe the universe, as it was devoid of matter, this did motivate astronomers to look for a relationship between distance and redshift. In 1924 Swedish astronomer Karl Lundmark published an empirical study that gave a roughly linear relation (though with lots of scatter) between the distances and velocities of the spirals. The difficulty was in knowing the distances accurately enough. Lundmark used novae that had been observed in the Andromeda Nebula to establish the distance of that nebula by assuming that these novae would have the same average absolute brightness as novae in the Milky Way whose distances were approximately known. For more-distant spirals, Lundmark invoked the crude assumptions that those spirals had to have the same diameter and brightness as the Andromeda Nebula. Thus, the novae functioned as standard candles (that is, objects with a defined brightness), and for more-distant spirals, the spirals themselves became the standard candle.

On the theoretical side, between 1922 and 1924 Russian mathematician Aleksandr Friedmann studied nonstatic cosmological solutions to Einstein’s equations. These went beyond Einstein’s model by allowing expansion or contraction of the universe and beyond de Sitter’s model by allowing the universe to contain matter. Friedmann also introduced cosmological models with negative curvature. (In a negatively curved space, the angles of a triangle add up to less than 180°.) Friedmann’s solutions had little immediate impact, partly because of his early death in 1925 and partly because he had not connected his theoretical work with astronomical observations. It did not help that Einstein published a note claiming that Friedmann’s 1922 paper contained a fundamental error; Einstein later withdrew this criticism.

The origin of the universe

Development of the big-bang theory

In 1927 Belgian physicist and cleric Georges Lemaître published a paper that put the theoretical and empirical squarely together under the title “Un Univers homogène de masse constante et de rayon croissant rendant compte de la vitesse radiale des nébuleuses extra-galactiques” (“A Homogeneous Universe of Constant Mass and Growing Radius, Accounting for the Radial Velocity of the Extragalactic Nebulae”). Lemaître began with a study of the dynamical solutions of Einstein’s model (with the cosmological constant included)—that is, those solutions with a cosmic radius that varies with time. He treated the Doppler shifts of the spiral nebulae as evidence of a cosmic expansion and used the redshifts and distances of 42 nebulae to deduce a value for the slope of the velocity-distance graph. At the time, Lemaître’s paper had little impact, partly because it had been published in the rather obscure Annales de la Societe Scientifique de Bruxelles (“Annals of the Scientific Society of Brussels”), and it was fully appreciated only a few years later, when cosmologists and astronomers had become more open to the idea of an expanding universe.

In 1929, building on Friedmann’s work, American mathematician and physicist Howard P. Robertson summarized the most general space-time metric that is possible under the assumption that the universe is homogeneous (of the same density everywhere) and isotropic (the same in all spatial directions). (A metric is a generalization of the Pythagorean theorem that describes the inherent geometry of space-time.) Similar results were obtained by English mathematician Arthur G. Walker, so this metric is called the Robertson-Walker metric. The Robertson-Walker metric and the expansion of the universe (as revealed by the galactic redshifts) were the twin foundations on which much of 20th-century cosmology was constructed.

American astronomer Edwin Hubble was the most influential observer of his generation. Using the 100-inch (254-cm) reflector at the Mount Wilson Observatory, in 1923 Hubble identified a Cepheid variable star in the Andromeda Nebula. From this he was able to determine a more-precise distance to the nebula, using the Cepheid variable as a much better standard candle. The Cepheids vary in brightness in a regular and easily identifiable way, with a quick increase in brightness followed by a slower decline. In 1908 American astronomer Henrietta Leavitt had found a relationship between the period and the brightness: the brighter the Cepheid, the longer its period. Ejnar Hertzsprung and American astronomer Harlow Shapley went on to calibrate the relationship in terms of absolute magnitudes. Hubble could easily measure the Cepheid’s period. He could then use the calibration curve to determine the star’s absolute magnitude, or intrinsic brightness, and the intrinsic brightness compared with the observed brightness gave the distance of the star. This measurement established beyond question that the Andromeda Nebula is outside the Milky Way and is a galaxy in its own right. Further work by Hubble with Cepheid variables in other spiral nebulae confirmed the island-universe theory.

When Hubble turned to the problem of the distance-redshift relationship, he soon superseded Slipher’s work. In 1929 Hubble published a paper showing a clear linear relationship between distance and redshift, which he interpreted as a velocity. He used Slipher’s velocities but added more that had been measured at Mount Wilson by American astronomer Milton Humason. Distances of the nearer nebulae were found by using Cepheids as standard candles. At greater distances Hubble used as a standard candle the brightest individual stars that could be resolved (assuming that these would be of the same brightness in all galaxies), and at greater distances yet, the luminosities of the nebulae themselves were the standard candle. Hubble’s paper led to a rapid acceptance of the distance-redshift (or distance-velocity) relation in the astronomical community, and this relationship is known as “Hubble’s law,” although, as discussed above, it had been several times anticipated.

Hubble himself was quite cautious about what the distance-velocity relationship implied about the history of the universe, but the natural conclusion to draw was that in the remote past all the galaxies had been close together. The distance-velocity relationship being linear, if galaxy B was 10 times farther away than galaxy A, it would be receding at 10 times the speed. By the same token, if the galactic clock was run backward to the beginning, both A and B would be at the same point (galaxy B retracing the greater distance at greater speed). Hubble’s value for the slope of the line in the velocity-versus-distance graph (today known as the Hubble constant) was 500 km (300 miles) per second per million parsecs (megaparsec). (A parsec is about 3.26 light-years and is the distance at which the radius of Earth’s orbit would subtend an angle of one second.) With this value for the Hubble constant, the universe appeared to be about two billion years old.

Subsequent studies indicated that this estimate was far too young. The study of radioactive isotopes in rocks suggested that Earth had to be 4.5 billion years old, which would make the universe younger than some of the objects in it. The value of the Hubble constant has been revised repeatedly. A major correction was made in 1952 when American astronomer Walter Baade discovered that Hubble had seriously underestimated galactic distances, because there are actually two different kinds of Cepheids. Baade’s recalibration resulted in a halving of the Hubble constant. A further major correction by American astronomer Allan Sandage in 1958 brought it down to about 100 km (60 miles) per second per megaparsec. Sandage, who was Hubble’s former observing assistant, showed that what Hubble had taken as the brightest individual stars in a galaxy were actually tight clusters of bright stars embedded in gaseous nebulae. For several decades the value of the constant was (according to different researchers) in the range 50–100 km (80–160 miles) per second per megaparsec. The currently accepted value for Hubble’s constant is around 71 km (44 miles) per second per megaparsec, with a margin of error of about 5 percent. The associated age of the universe, tightly constrained by many types of observations, is about 13.7 billion years.

Several astronomers proposed mechanisms to explain the redshifts without accepting the expansion of the universe. In 1929 the Swiss astrophysicist Fritz Zwicky proposed that photons gradually give up their energy to the intergalactic matter through which they travel, through a process analogous to Compton scattering, leading to a progressive reddening of the light. Others simply suggested various versions of the reddening of light with distance (collectively these were called the “tired light” hypothesis) without attempting to provide a physical explanation. These proposals never commanded a wide following, and during the 1930s astronomers and cosmologists increasingly embraced the expansion of the universe.

The general-relativistic cosmological models and the observed expansion of the universe suggest that the universe was once very small. In the 1930s astronomers began to explore evolutionary models of the universe, a good example being Georges Lemaître’s primeval atom. According to Lemaître, the universe began as a single atom having an atomic weight equal to the entire mass of the universe, which then decayed by a super-radiative process until atoms of ordinary atomic weight emerged.

A pioneering study of elemental abundances in the stars had been made by British-born American astronomer Cecilia Payne in her doctoral thesis of 1925. The amount of each element present in a star can be inferred from the strengths of the absorption lines in the star’s spectrum, if these are controlled for the temperature and pressure of the star. One fact that emerged early on was that stars did not have the same composition as Earth and were predominantly hydrogen and helium. In 1938 Norwegian mineralogist Victor Goldschmidt published a detailed summary of data on cosmic abundances of the elements, running over most of the periodic table.

Although it is possible to see Lemaître’s theory as a progenitor of the “big bang” theory, it was a paper of 1948 by American physicist Ralph Alpher and his dissertation supervisor, George Gamow, that changed the direction of research by putting nuclear physics into cosmology. As a joke, Gamow added the name of physicist Hans Bethe in order to preserve the Alpher-Bethe-Gamow sequence of (almost) Greek letters. In the aßγ paper, which was only one page long, Alpher and Gamow maintained that the formation of the elements (nucleosynthesis) began about 20 seconds after the start of the expansion of the universe. They supposed that the universe began with a hot dense gas of neutrons, which started to decay into protons and electrons. The building up of the elements was due to successive neutron capture (and readjustments of charge by ß-decay). Using recently published values for the neutron-capture cross-sections of the elements, they integrated their equations to produce a graph of the abundances of all the elements, which resulted in a smooth-curve approximation to the jagged abundance curve that had been published by Goldschmidt.

In another paper in 1948, Alpher and American physicist Robert Herman argued that electromagnetic radiation from the early universe should still exist, but with the expansion it should now correspond to a temperature of about 5 K (kelvins, or −268 °C [−451 °F]) and thus would be visible to radio telescopes. In a 1953 paper, Alpher, Herman, and American physicist James Follin provided a stage-by-stage history of the early universe, concluding that nucleosynthesis was essentially complete after 30 minutes of cosmic expansion. They deduced that if all the neutrons available at the end of nucleosynthesis went into making helium only, the present-day hydrogen-to-helium ratio would be between 7:1 and 10:1 in terms of numbers of atoms. This would correspond to a present-day universe that was between 29 and 36 percent helium by weight. (Because some neutrons would go into building other elements, the helium figures would be upper limits.) They pointed out that these figures were of the same order as the hydrogen-to-helium ratios measured in planetary nebulae and stellar atmospheres, though these showed quite a large range.

The Gamow-Alpher theory largely ceased development after 1953, and it failed to attract a following, in spite of the fact that they had published in highly prominent journals and had made detailed, testable predictions. Unfortunately, it was not until the 1960s that the hydrogen-to-helium ratio became known precisely enough to test the theory. More crucially, Alpher and Gamow failed to interest radio astronomers in looking for the 5-K background radiation, and their prediction was soon forgotten.

The steady-state challenge

In England, also in 1948, an alternative theory emerged called the steady-state universe. Different versions of it were proposed by English mathematician and astronomer Fred Hoyle and by the team of British mathematician and cosmologist Hermann Bondi and British astronomer Thomas Gold, but the key idea was that although the universe was expanding, its average properties did not change with time. As the universe expanded, the density of matter would be expected to diminish, but new hydrogen atoms were created that formed clouds of gas that condensed into new stars and galaxies. The number of new hydrogen atoms required per year was so tiny that one could not hope to observe this process directly. However, there were predictable observational consequences that should allow one to distinguish between a steady-state universe or a big-bang universe. (The term big bang was coined by Hoyle as a mildly pejorative characterization of the rival theory in a radio talk in 1949.)

For example, in a big-bang universe, when one looks at galaxies that are far away, one also sees them as they were in the remote past (because of the travel time of the light). Thus, one might expect that distant galaxies are less-evolved or that they contain more young stars. But in a steady-state universe, one would see galaxies at all possible stages of evolutionary development at even the farthest distances. The density of galaxies in space should also diminish with time in a big-bang universe. Therefore, galaxies at great distances should be more densely crowded together than nearby galaxies are. But in a steady-state universe, the average density of galaxies should be about the same everywhere and at every time. In the 1950s the Cambridge radio astronomer Martin Ryle showed that there were more radio galaxies at great distances than there were nearby, thus showing that the universe had evolved over time, a result that could not be explained in steady-state theory.

The discovery of quasars (quasi-stellar radio sources) in the early 1960s also told heavily against the steady-state theory. Quasars were first identified as strong radio sources that in visible light appear to be identified with small starlike objects. Further, they have large redshifts, which implies that they are very far away. From their distance and their apparent luminosity, it was inferred that they emit copious amounts of energy; a single quasar might be brighter than a whole galaxy. There was no room for such objects in a steady-state universe, in which the contents of any region of space (seen as it is now or as it was long ago) should be roughly similar. The quasars were a clear sign that the universe was evolving.

Steady-state theory never had a large following, and its supporters were centred in Britain. Nevertheless, having a competing theory forced the big-bang cosmologists to strengthen their arguments and to collect supporting data. A key question centred on the abundances and origins of the chemical elements. In steady-state theory, it was essential that all the elements could be synthesized in stars. By contrast, in the aßγ paper, Alpher and Gamow tried to show that all the elements could be made in the big bang. Of course, in a more reasonable view, big-bang theorists had to accept that some element formation does take place in stars, but they were keen to show that the stars could not account for all of it. In particular, the stars could not be the source of most of the light elements. For example, it was impossible to see how during the lifetime of a galaxy the stars could build up the helium content to 30 percent.

One obstacle for big-bang theory was the absence of any stable isotopes at atomic mass 5 or 8. In 1952 Austrian-born American astrophysicist Edwin Salpeter proposed that three alpha particles (helium nuclei) can come together to produce carbon-12 and that this happens often enough to resolve the mass-gap problem in the interiors of stars. However, conditions in the early universe were not right for bridging the mass gap in this way, so the mass-gap problem was seen as favouring steady-state theory. Hoyle adopted Salpeter’s proposal in 1953. In 1957 Hoyle, with American astronomers William Fowler, Margaret Burbidge, and Geoffrey Burbidge (or B2FH, as their paper was later called), gave an impressive and detailed account of the abundances of most elements in terms of conditions appropriate to stellar interiors. Although the B2FH paper was not explicitly a steady-state theory, it was often seen as favouring that model, as it had not made use of temperature and pressure conditions appropriate to the big bang. But in papers of 1964 (with English astrophysicist Roger Tayler) and 1967 (with Fowler and American physicist Robert Wagoner), Hoyle concluded that the lighter elements could be built up satisfactorily only in conditions like those of the big bang. Hoyle himself continued to favour supermassive objects as the origin of the elements over the big bang, but most astronomers saw this work as vindicating big-bang theory. In defending a failed cosmological theory, Hoyle had done an enormous amount of good work of lasting value on nucleosynthesis.

When good estimates of the cosmic abundance of deuterium and other light elements became available, big-bang theory proved capable of detailed explanation of the cosmic abundances of all the light elements. In current scenarios, hydrogen (H) and its heavier isotope, deuterium (2H), most of the two helium isotopes (3He and 4He), and lithium (7Li) were produced shortly after the big bang. Given whatever one assumes about the present-day density of matter in the universe, one can calculate what sort of cosmic abundances should have resulted from the big bang. It is regarded as a triumph of the big-bang model that the present-day abundances of these elements can all be explained from one set of initial conditions. According to current thinking, most of the heavier elements were then built up in stars, neutron star mergers, and supernova explosions.

The cosmic microwave background proves the theory

In 1965 American astronomers Arno Penzias and Robert W. Wilson were working at Bell Laboratories on a 6-metre (20-foot) horn antenna. The original purpose of the antenna was to detect reflected signals from high-altitude balloons, with the goal of applying the technology to communications satellites, but Penzias and Wilson had adapted it for doing radio astronomy. They detected a constant, persistent signal, corresponding to an excess temperature of 3.3 K (−269.9 °C [−453.7 °F]). After eliminating every source of circuit noise they could think of, and even shooing a pair of pigeons that had been roosting (and leaving behind “white dielectric material”) in the horn, they found that the signal remained and that it was constant, no matter in which direction the telescope was pointed. At nearby Princeton University, they consulted with American physicist Robert Dicke, who was studying oscillatory models of the universe with hot phases and who was therefore not surprised by what they had found. About the same time, astrophysicist James Peebles, Dicke’s former student, also published a paper predicting the existence of a universal background radiation at a temperature of 10 K (−263 °C [−441 °F]), apparently completely unaware of Alpher and Herman’s earlier prediction. Suddenly the pieces fell together. The cosmic microwave background (CMB) was accepted as the third major piece of evidence in support of the big-bang theory. In the early stages of the expansion, when atoms were all still completely ionized, the universe was opaque to electromagnetic radiation. But when the universe cooled enough to allow the formation of neutral atoms, it suddenly became transparent to electromagnetic radiation (just as light can travel through air). At this “decoupling time,” the electromagnetic radiation was of very high energy and very short wavelengths. With the continued expansion of space, wavelengths were stretched until they reached their current microwave lengths (from about a millimetre to tens of centimetres in wavelength). Thus, every bit of empty space acts as a source of radio waves—a phenomenon predicted (twice!) by big-bang theory but for which steady-state theory had no ready explanation. For most cosmologists, this marked the end of the steady-state theory, even though Hoyle and his collaborators continued to tweak and adjust the theory to try to meet objections.

By the mid-1960s, big-bang theory had become the standard cosmology, underpinned by the observed expansion, the measured abundances of the light elements, and the presence of the cosmic microwave background. Of course, the theory was eventually to acquire many different forms and refinements.

Echoes of the big bang

Dark matter

Over the course of the 20th century, it became clear that there is much more to the universe than meets the eye. On the basis of early estimates of the mass density of the Milky Way, English physicist and mathematician James Jeans suggested in 1922 that the galaxy might contain three times as many dark stars as visible ones. In 1933 Fritz Zwicky, by studying the dynamics of clusters of galaxies, concluded that there is not enough visible matter in the galaxies to hold the clusters together gravitationally. He also pointed out that the measured quantity of luminous matter was far below the value that would be necessary for critical density—i.e., to produce a universe with an expansion that would gradually slow to a halt at infinity—but he speculated that the dark matter could conceivably be enough to make up the difference.

Jeans’s and Zwicky’s comments did not attract a lot of attention, and dark matter became a central issue only in the 1970s. In 1974 Peebles, Jeremiah Ostriker, and Amos Yahil in the United States and Jaan Einasto, Ants Kaasik, and Enn Saar in Soviet Estonia concluded, on the basis of studies of galactic dynamics, that 90–95 percent of the universe must be in the form of dark matter. American astronomer Vera Rubin published a paper in 1978 studying the rotational velocities of stars in galaxies as a function of their distances from the galactic centre. Rotational velocities were found to be nearly constant over a fairly large radial distance, though predictions based on the distribution of visible matter implied that they would decrease with distance. Rubin’s discoveries were interpreted as evidence for the presence of substantial amounts of dark matter in the haloes around galaxies. About the same time, radio astronomers, using a spectral line of hydrogen at 21-cm wavelength, obtained a similar result in the outer parts of galaxies where there is little starlight. Present-day thinking is that the universe is very close to flat (Euclidean) in its geometry, which implies that it is close to critical density. However, the nucleosynthesis calculations show agreement with the present-day abundances of the light elements only if one supposes that ordinary baryonic matter (i.e., matter made of protons and neutrons) accounts for no more than about 5 percent of the critical density.

Candidates for dark matter in the form of ordinary baryonic matter include black holes, Jupiter-sized planets, and brown dwarfs (starlike objects that are too small to ignite nuclear reactions in their interiors). Some of the new grand unified theories (GUTs) of particle physics predict the existence of large quantities of exotic fundamental particles, called weakly interacting massive particles (WIMPs). The 1998 discovery that neutrinos have mass (they had been considered perfectly massless since Austrian-born physicist Wolfgang Pauli’s prediction of them in 1930) provides a small part of the answer. But the nature of the bulk of dark matter is still unknown.

Satellite observatories

By placing astronomical instruments in space, they would be free from the interference of Earth’s atmosphere. Observing instruments in space have played important roles since the age of artificial satellites began with Sputnik in 1957. Astronomical instruments had earlier been sent aloft on balloons and rockets, but satellites permitted vastly longer observing times and greater stability. The very first U.S. satellite, Explorer 1, launched in 1958 as a project designed for the International Geophysical Year, was involved in a major discovery. The radiation detector on board gave the first signs of the belts of energetic charged particles that surround Earth (the Van Allen belts, named for American physicist James Van Allen). Beginning in 1962, a series of eight Orbital Solar Observatories monitored the Sun for more than a complete sunspot cycle and had far clearer views of the Sun’s corona than could be obtained from Earth-based observatories, because of the distortion of optical images by Earth’s atmosphere.

The first successful planetary flyby was that of Venus in 1962 by Mariner 2, which carried several instruments but no cameras. The first flyby to return images was the Mariner 4 mission in 1965, which sent back 22 images of Mars. The first flybys of Jupiter and SaturnPioneer 10 (1973) and Pioneer 11 (1979)—sent back spectacular images of the planets and their rings and satellites that fundamentally altered planetary science and captured the public imagination. Specialized satellites have extended astronomical observing into the infrared, gamma-ray, and X-ray portions of the spectrum.

In 1989 the Cosmic Background Explorer (COBE) satellite began precise measurements of the microwave background radiation. This gave, by 1994, a perfect fit to a blackbody spectrum corresponding to 2.726K (−270.424 °C [−454.763 °F]). However, the most significant result, announced by American physicist George Smoot in 1992, was COBE’s detection of small fluctuations in the temperature in different directions in space—variations as small as a few parts in 100,000—that correspond to density fluctuations in the early universe at the decoupling time, about 300,000 years after the big bang. This discovery came as a relief to cosmologists, because the earlier failure to detect fluctuations in the spectrum was starting to cause difficulties for theories of structure formation in the early universe.

By far the most ambitious instrument put into Earth orbit was the Hubble Space Telescope (HST), launched in 1990. Shortly afterward it was discovered that a design flaw in the principal mirror greatly reduced the image quality, but this was fixed by compensating optical devices inserted on a subsequent service trip by astronauts to the telescope. Among the original missions of the HST were determining more accurate values of the Hubble constant and the deceleration parameter, with the goal of limiting the number of possible cosmological models. The deceleration parameter is a measure of the rate at which the expansion of the universe is slowing down as the universe expands against gravity.

Dark energy

In the 1980s astronomers began to use Type Ia supernovae as standard candles. These are believed to come about in the following way. A white dwarf star in a binary orbit with a neighbour can slowly pull material off, gradually increasing its own mass. Ordinarily the mass of the white dwarf could not exceed the Chandrasekhar limit of about 1.4 solar masses, or it would collapse to form a neutron star. However, in the case of white dwarfs rich in carbon, with the slow accretion of material pulled from the neighbour, the core temperature rises until the nuclear ignition of carbon causes a runaway explosion. Because of the slow accretion and the mass limit, these supernovae are remarkably uniform in their brightness; moreover, because they are so bright, they can be seen at great distances. In short, the uniform and extreme brightness of Type Ia supernovae make them excellent standard candles.

In the 1990s two groups used observations of Type Ia supernovae in distant galaxies to work out distances to those galaxies, and thus how the rate of the universe’s expansion changed over time, more precisely than ever before. The Supernova Cosmology Project, led by American physicist Saul Perlmutter, and the High-Z Supernova Search Team, directed by Australian astronomer Brian Schmidt and American astronomer Adam Riess, used observations taken with ground-based telescopes as well as with the HST. The result was most unexpected. Far from finding a better value for the deceleration parameter, after a period of confusion and contradiction, both groups found that the expansion of the universe is actually speeding up. The direct observations were that distant supernovae appeared to be 20–25 percent dimmer than expected. The two teams ruled out such possibilities as dimming by dust, and their papers, published in 1998 and 1999, led to the same general conclusion. The expansion of the universe is accelerating, and that acceleration began only about five billion or six billion years ago.

The consensus emerging from the Ia supernovae projects was that the geometry of the universe is essentially flat, and therefore quite close to the critical density, with matter making up only about 30 percent of the total energy density and “dark energy” making up the remaining 70 percent. (Subsequent research has slightly modified these figures.) Although other possibilities are open, the dark energy is often identified with an Einsteinian cosmological constant that provides a universal repulsive force, which explains the acceleration. The nature of the dark energy is unknown. It may be connected with quantum-mechanical vacuum energy; however, there are serious unresolved difficulties with this possibility. Of the roughly 30 percent of the universe that is matter, only about 5 percent can be ordinary baryonic matter. Of this, only a small part is visible in the form of planets, stars, and galaxies.

The objects of all astronomical inquiry, from the time of the ancient Greeks and Babylonians to the 20th century, thus represent only the tip of the iceberg. After almost 4,000 years of astronomy, the universe is no less strange than it must have seemed to the Babylonians. James Evans



Return to Top

Copyright
Photos: Encyclopedia Britannica & Respective Authors
Illustration: Encyclopedia Britannica & Respective Authors
Source: Encyclopedia Britannica

This web site is a Jan.-Feb. 2018 accessed archive copy of the original version, saved at TYCHOS.info to preserve the web reference. External links may no longer be valid.

Source URL: https://www.britannica.com/science/astronomy/The-techniques-of-astronomy

According to the “Fair Use” clause of International Copyright Law, the authors declare that the use of the photos, videos and information in this academic research are analyzed for purposes of “criticism, comment, news reporting, teaching, scholarship, or research” according to Section 107 of Title 17 of the US Code.

- Limitations on exclusive rights: Fair use, U.S. Copyright Code.