geometry

geometry
/jee om"i tree/, n.
1. the branch of mathematics that deals with the deduction of the properties, measurement, and relationships of points, lines, angles, and figures in space from their defining conditions by means of certain assumed properties of space.
2. any specific system of this that operates in accordance with a specific set of assumptions: Euclidean geometry.
3. the study of this branch of mathematics.
4. a book on this study, esp. a textbook.
5. the shape or form of a surface or solid.
6. a design or arrangement of objects in simple rectilinear or curvilinear form.
[1300-50; ME < L geometria < Gk geometría. See GEO-, -METRY]

* * *

(as used in expressions)
non Euclidean geometry

* * *

Introduction

      the branch of mathematics concerned with the shape of individual objects, spatial relationships among various objects, and the properties of surrounding space. It is one of the oldest branches of mathematics, having arisen in response to such practical problems as those found in surveying, and its name is derived from Greek words meaning “Earth measurement.” Eventually it was realized that geometry need not be limited to the study of flat surfaces (plane geometry) and rigid three-dimensional objects (solid geometry) but that even the most abstract thoughts and images might be represented and developed in geometric terms.

      This article begins with a brief guidepost to the major branches of geometry and then proceeds to an extensive historical treatment. For information on specific branches of geometry, see Euclidean geometry, analytic geometry, projective geometry, differential geometry, non-Euclidean geometries (non-Euclidean geometry), and topology.

Major branches of geometry

      In several ancient cultures there developed a form of geometry suited to the relationships between lengths, areas, and volumes of physical objects. This geometry was codified in Euclid's Elements about 300 BC on the basis of 10 axioms, or postulates, from which several hundred theorems were proved by deductive logic. The Elements epitomized the axiomatic-deductive method for many centuries.

      Analytic geometry was initiated by the French mathematician René Descartes (Descartes, René) (1596–1650), who introduced rectangular coordinates to locate points and to enable lines and curves to be represented with algebraic equations. Algebraic geometry is a modern extension of the subject to multidimensional and non-Euclidean spaces.

      Projective geometry originated with the French mathematician Girard Desargues (Desargues, Girard) (1591–1661) to deal with those properties of geometric figures that are not altered by projecting their image, or “shadow,” onto another surface.

      The German mathematician Carl Friedrich Gauss (Gauss, Carl Friedrich) (1777–1855), in connection with practical problems of surveying and geodesy, initiated the field of differential geometry. Using differential calculus, he characterized the intrinsic properties of curves and surfaces. For instance, he showed that the intrinsic curvature of a cylinder is the same as that of a plane, as can be seen by cutting a cylinder along its axis and flattening, but not the same as that of a sphere, which cannot be flattened without distortion.

Non-Euclidean geometries (non-Euclidean geometry)
      Beginning in the 19th century, various mathematicians substituted alternatives to Euclid's parallel postulate, which, in its modern form, reads, “given a line and a point not on the line, it is possible to draw exactly one line through the given point parallel to the line.” They hoped to show that the alternatives were logically impossible. Instead, they discovered that consistent non-Euclidean geometries exist.

Topology
      Topology, the youngest and most sophisticated branch of geometry, focuses on the properties of geometric objects that remain unchanged upon continuous deformation—shrinking, stretching, and folding, but not tearing. The continuous development of topology dates from 1911, when the Dutch mathematician L.E.J. Brouwer (Brouwer, Luitzen Egbertus Jan) (1881–1966) introduced methods generally applicable to the topic.

History of geometry
 The earliest known unambiguous examples of written records—dating from Egypt and Mesopotamia about 3100 BC—demonstrate that ancient peoples had already begun to devise mathematical rules and techniques useful for surveying land areas, constructing buildings, and measuring storage containers. Beginning about the 6th century BC, the Greeks gathered and extended this practical knowledge and from it generalized the abstract subject now known as geometry, from the combination of the Greek words geo (“Earth”) and metron (“measure”) for the measurement of the Earth.

      In addition to describing some of the achievements of the ancient Greeks, notably Euclid's logical development of geometry in the Elements, this article examines some applications of geometry to astronomy, cartography, and painting from classical Greece through medieval Islam and Renaissance Europe. It concludes with a brief discussion of extensions to non-Euclidean and multidimensional geometries in the modern age.

Ancient geometry: practical and empirical
      The origin of geometry lies in the concerns of everyday life. The traditional account, preserved in Herodotus's History (5th century BC), credits the Egyptians with inventing surveying in order to reestablish property values after the annual flood of the Nile. Similarly, eagerness to know the volumes of solid figures derived from the need to evaluate tribute, store oil and grain, and build dams and pyramids. Even the three abstruse geometrical problems of ancient times—to double a cube, trisect an angle, and square a circle, all of which will be discussed later—probably arose from practical matters, from religious ritual, timekeeping, and construction, respectively, in pre-Greek societies of the Mediterranean. And the main subject of later Greek geometry, the theory of conic sections (conic section), owed its general importance, and perhaps also its origin, to its application to optics and astronomy.

      While many ancient individuals, known and unknown, contributed to the subject, none equaled the impact of Euclid and his Elements of geometry, a book now 2,300 years old and the object of as much painful and painstaking study as the Bible. Much less is known about Euclid, however, than about Moses. In fact, the only thing known with a fair degree of confidence is that Euclid taught at the Library of Alexandria (Alexandria, Library of) during the reign of Ptolemy I (323–285/83 BC). Euclid wrote not only on geometry but also on astronomy and optics and perhaps also on mechanics and music. Only the Elements, which was extensively copied and translated, has survived intact.

      Euclid's Elements was so complete and clearly written that it literally obliterated the work of his predecessors. What is known about Greek geometry before him comes primarily from bits quoted by Plato and Aristotle and by later mathematicians and commentators. Among other precious items they preserved are some results and the general approach of Pythagoras (c. 580–c. 500 BC) and his followers. The Pythagoreans (Pythagoreanism) convinced themselves that all things are, or owe their relationships to, numbers. The doctrine gave mathematics supreme importance in the investigation and understanding of the world. Plato developed a similar view, and philosophers influenced by Pythagoras or Plato often wrote ecstatically about geometry as the key to the interpretation of the universe. Thus ancient geometry gained an association with the sublime to complement its earthy origins and its reputation as the exemplar of precise reasoning.

Finding the right angle
 Ancient builders and surveyors needed to be able to construct right angles in the field on demand. The method employed by the Egyptians earned them the name “rope pullers” in Greece, apparently because they employed a rope for laying out their construction guidelines. One way that they could have employed a rope to construct right triangles was to mark a looped rope with knots so that, when held at the knots and pulled tight, the rope must form a right triangle. The simplest way to perform the trick is to take a rope that is 12 units long, make a knot 3 units from one end and another 5 units from the other end, and then knot the ends together to form a loop, as shown in the animation—>. However, the Egyptian scribes have not left us instructions about these procedures, much less any hint that they knew how to generalize them to obtain the Pythagorean theorem: the square on the line opposite the right angle equals the sum of the squares on the other two sides. Similarly, the Vedic scriptures of ancient India contain sections called sulvasutras, or “rules of the rope,” for the exact positioning of sacrificial altars. The required right angles were made by ropes marked to give the triads (3, 4, 5) and (5, 12, 13).

      In Babylonian clay tablets (c. 1700–1500 BC) modern historians have discovered problems whose solutions indicate that the Pythagorean theorem and some special triads were known more than a thousand years before Euclid. A right triangle made at random, however, is very unlikely to have all its sides measurable by the same unit—that is, every side a whole-number multiple of some common unit of measurement. This fact, which came as a shock when discovered by the Pythagoreans, gave rise to the concept and theory of incommensurability.

Locating the inaccessible
  By ancient tradition, Thales of Miletus, who lived before Pythagoras in the 6th century BC, invented a way to measure inaccessible heights, such as the Egyptian pyramids. Although none of his writings survives, Thales may well have known about a Babylonian observation that for similar triangles (triangles having the same shape but not necessarily the same size) the length of each corresponding side is increased (or decreased) by the same multiple. A determination of the height of a tower using similar triangles is demonstrated in the figure—>. The ancient Chinese arrived at measures of inaccessible heights and distances by another route, using “complementary” rectangles, as seen in the next figure—>, which can be shown to give results equivalent to those of the Greek method involving triangles.

Estimating the wealth
      A Babylonian cuneiform tablet written some 3,500 years ago treats problems about dams, wells, water clocks, and excavations. It also has an exercise on circular enclosures with an implied value of π = 3. The contractor for King Solomon's swimming pool, who made a pond 10 cubits across and 30 cubits around (1 Kings 7:23), used the same value. However, the Hebrews should have taken their π from the Egyptians before crossing the Red Sea, for the Rhind papyrus (c. 2000 BC; our principal source for ancient Egyptian mathematics) implies π = 3.1605.

      Knowledge of the area of a circle was of practical value to the officials who kept track of the pharaoh's tribute as well as to the builders of altars and swimming pools. Ahmes, the scribe who copied and annotated the Rhind papyrus (c. 1650 BC), has much to say about cylindrical granaries and pyramids, whole and truncated. He could calculate their volumes, and, as appears from his taking the Egyptian seked, the horizontal distance associated with a vertical rise of one cubit, as the defining quantity for the pyramid's slope, he knew something about similar triangles.

Ancient geometry: abstract and applied
The three classical problems
      In addition to proving mathematical theorems, ancient mathematicians constructed various geometrical objects. Euclid arbitrarily restricted the tools of construction to a straightedge (an unmarked ruler) and a compass. The restriction made three problems of particular interest (to double a cube, to trisect an arbitrary angle, and to square a circle) very difficult—in fact, impossible. Various methods of construction using other means were devised in the classical period, and efforts, always unsuccessful, using straightedge and compass persisted for the next 2,000 years. In 1837 the French mathematician Pierre Laurent Wantzel proved that doubling the cube and trisecting the angle are impossible, and in 1880 the German mathematician Ferdinand von Lindemann (Lindemann, Ferdinand von) showed that squaring the circle is impossible, as a consequence of his proof that π is a transcendental number (pi).

Doubling the cube
      The Vedic scriptures made the cube the most advisable form of altar for anyone who wanted to supplicate in the same place twice. The rules of ritual required that the altar for the second plea have the same shape but twice the volume of the first. If the sides of the original and derived altars are a and b, respectively, then b3 = 2a3. The problem came to the Greeks together with its ceremonial content. An oracle disclosed that the citizens of Delos could free themselves of a plague merely by replacing an existing altar by one twice its size. The Delians applied to Plato. He replied that the oracle did not mean that the gods wanted a larger altar but that they had intended “to shame the Greeks for their neglect of mathematics and their contempt for geometry.” With this blend of Vedic practice, Greek myth, and academic manipulation, the problem of the duplication of the cube took a leading place in the formation of Greek geometry.

       Hippocrates of Chios, who wrote an early Elements about 450 BC, took the first steps in cracking the altar problem. He reduced the duplication to finding two mean proportionals between 1 and 2, that is, to finding lines x and y in the ratio 1:x = x:y = y:2. After the intervention of the Delian oracle, several geometers around Plato's Academy found complicated ways of generating mean proportionals.

      A few generations later, Eratosthenes of Cyrene (c. 276–c. 194 BC) devised a simple instrument with moving parts that could produce approximate mean proportionals.

Trisecting the angle
      The Egyptians told time at night by the rising of 12 asterisms (constellations), each requiring on average two hours to rise. In order to obtain more convenient intervals, the Egyptians subdivided each of their asterisms into three parts, or decans. That presented the problem of trisection. It is not known whether the second celebrated problem of archaic Greek geometry, the trisection of any given angle, arose from the difficulty of the decan, but it is likely that it came from some problem in angular measure.

      Several geometers of Plato's time tried their hands at trisection. Although no one succeeded in finding a solution with straightedge and compass, they did succeed with a mechanical device and by a trick. The mechanical device, perhaps never built, creates what the ancient geometers called a quadratrix. Invented by a geometer known as Hippias Of Elis (fl. 5th century BC), the quadratrix is a curve traced by the point of intersection between two moving lines, one rotating uniformly through a right angle, the other gliding uniformly parallel to itself. (See Sidebar: Trisecting the Angle: The Quadratrix of Hippias.)

      The trick for trisection is an application of what the Greeks called neusis, a maneuvering of a measured length into a special position to complete a geometrical figure. A late version of its use, ascribed to Archimedes (c. 285–212/211 BC), exemplifies the method of angle trisection. (See Sidebar: Trisecting the Angle: Archimedes' Method.)

Squaring the circle
      The pre-Euclidean Greek geometers transformed the practical problem of determining the area of a circle into a tool of discovery. Three approaches can be distinguished: Hippocrates' dodge of substituting one problem for another; the application of a mechanical instrument, as in Hippias's device for trisecting the angle; and the technique that proved the most fruitful, the closer and closer approximation to an unknown magnitude difficult to study (e.g., the area of a circle) by a series of known magnitudes easier to study (e.g., areas of polygons)—a technique known in modern times as the “method of exhaustion (exhaustion, method of)” and attributed by its greatest practitioner, Archimedes, to Plato's student Eudoxus of Cnidus (c. 408–c. 355 BC).

      While not able to square the circle, Hippocrates did demonstrate the quadratures of lunes; that is, he showed that the area between two intersecting circular arcs could be expressed exactly as a rectilinear area and so raised the expectation that the circle itself could be treated similarly. (See Sidebar: Quadrature of the Lune.) A contemporary of Hippias's discovered that the quadratrix could be used to almost rectify circles. These were the substitution and mechanical approaches.

 The method of exhaustion as developed by Eudoxus approximates a curve or surface by using polygons with calculable perimeters and areas. As the number of sides of a regular polygon inscribed in a circle increases indefinitely, its perimeter and area “exhaust,” or take up, the circumference and area of the circle to within any assignable error of length or area, however small. In Archimedes' usage, the method of exhaustion produced upper and lower bounds for the value of π, the ratio of any circle's circumference to its diameter. This he accomplished by inscribing a polygon within a circle, and circumscribing a polygon around it as well, thereby bounding the circle's circumference between the polygons' calculable perimeters. He used polygons with 96 sides and thus bound π between 310/71 and 31/7. (See animation—>.)

Idealization and proof
      The last great Platonist and Euclidean commentator of antiquity, Proclus (c. AD 410–485), attributed to the inexhaustible Thales the discovery of the far-from-obvious proposition that even apparently obvious propositions need proof. Proclus referred especially to the theorem, known in the Middle Ages as the , that in an isosceles triangle the angles opposite the equal sides are equal. The theorem may have earned its nickname from the Euclidean figure or from the commonsense notion that only an ass would require proof of so obvious a statement. (See Sidebar: The Bridge of Asses.)

      The ancient Greek geometers soon followed Thales over the Bridge of Asses. In the 5th century BC the philosopher-mathematician Democritus (c. 460–c. 370 BC) declared that his geometry excelled all the knowledge of the Egyptian rope pullers because he could prove what he claimed. By the time of Plato, geometers customarily proved their propositions. Their compulsion and the multiplication of theorems it produced fit perfectly with the endless questioning of Socrates and the uncompromising logic of Aristotle. Perhaps the origin, and certainly the exercise, of the peculiarly Greek method of mathematical proof should be sought in the same social setting that gave rise to the practice of philosophy—that is, the Greek polis. There citizens learned the skills of a governing class, and the wealthier among them enjoyed the leisure to engage their minds as they pleased, however useless the result, while slaves attended to the necessities of life. Greek society could support the transformation of geometry from a practical art to a deductive science. Despite its rigour, however, Greek geometry does not satisfy the demands of the modern systematist. Euclid himself sometimes appeals to inferences drawn from an intuitive grasp of concepts such as point and line or inside and outside, uses superposition, and so on. It took more than 2,000 years to purge the Elements of what pure deductivists deemed imperfections.

The Euclidean synthesis
      Euclid, in keeping with the self-conscious logic of Aristotle, began the first of his 13 books of the Elements with sets of definitions (“a line is breadthless length”), common notions (“the whole is greater than the part”), and axioms, or postulates (“all right angles are equal”). Of this preliminary matter, the fifth and last postulate, which states a sufficient condition that two straight lines meet if sufficiently extended, has received by far the greatest attention. In effect it defines parallelism. Many later geometers tried to prove the fifth postulate using other parts of the Elements. Euclid saw farther, for coherent geometries (known as non-Euclidean geometries (non-Euclidean geometry)) can be produced by replacing the fifth postulate with other postulates that contradict Euclid's choice.

      The first six books contain most of what Euclid delivers about plane geometry. Book I presents many propositions doubtless discovered by his predecessors, from Thales' equality of the angles opposite the equal sides of an isosceles triangle to the Pythagorean theorem, with which the book effectively ends. (See Sidebar: Euclid's Windmill.)

      Book VI applies the theory of proportion from Book V to similar figures and presents the geometrical solution to quadratic equations. As usual, some of it is older than Euclid. Books VII–X, which concern various sorts of numbers, especially primes, and various sorts of ratios, are seldom studied now, despite the importance of the masterful Book X, with its elaborate classification of incommensurable magnitudes, to the later development of Greek geometry. (See Sidebar: Incommensurables.)

      Books XI–XIII deal with solids: XI contains theorems about the intersection of planes and of lines and planes and theorems about the volumes of parallelepipeds (solids with parallel parallelograms as opposite faces); XII applies the method of exhaustion introduced by Eudoxus to the volumes of solid figures, including the sphere; XIII, a three-dimensional analogue to Book IV, describes the Platonic solids (Platonic solid). Among the jewels in Book XII is a proof of the recipe used by the Egyptians for the volume of a pyramid.

Gnomonics and the cone
      During its daily course above the horizon the Sun appears to describe a circular arc. Supplying in his mind's eye the missing portion of the daily circle, the Greek astronomer could imagine that his real eye was at the apex of a cone, the surface of which was defined by the Sun's rays at different times of the day and the base of which was defined by the Sun's apparent diurnal course. Our astronomer, using the pointer of a sundial, known as a gnomon, as his eye, would generate a second, shadow cone spreading downward. The intersection of this second cone with a horizontal surface, such as the face of a sundial, would give the trace of the Sun's image (or shadow) during the day as a plane section of a cone. (The possible intersections of a plane with a cone, known as the conic sections (conic section), are the circle, ellipse, point, straight line, parabola, and hyperbola.)

      However, the doxographers ascribe the discovery of conic sections to a student of Eudoxus's, Menaechmus (mid-4th century BC), who used them to solve the problem of duplicating the cube. His restricted approach to conics—he worked with only right circular cones and made his sections at right angles to one of the straight lines composing their surfaces—was standard down to Archimedes' era. Euclid adopted Menaechmus's approach in his lost book on conics, and Archimedes followed suit. Doubtless, however, both knew that all the conics can be obtained from the same right cone by allowing the section at any angle.

      The reason that Euclid's treatise on conics perished is that Apollonius of Perga (c. 262–c. 190 BC) did to it what Euclid had done to the geometry of Plato's time. Apollonius reproduced known results much more generally and discovered many new properties of the figures. He first proved that all conics are sections of any circular cone, right or oblique. Apollonius introduced the terms ellipse, hyperbola, and parabola for curves produced by intersecting a circular cone with a plane at an angle less than, greater than, and equal to, respectively, the opening angle of the cone.

Astronomy and trigonometry

      In an inspired use of their geometry, the Greeks did what no earlier people seems to have done: they geometrized the heavens by supposing that the Sun, Moon, and planets move around a stationary Earth on a rotating circle or set of circles, and they calculated the speed of rotation of these supposititious circles from observed motions. Thus they assigned to the Sun a circle eccentric to the Earth to account for the unequal lengths of the seasons.

       Ptolemy (fl. AD 127–145 in Alexandria, Egypt) worked out complete sets of circles for all the planets. In order to account for phenomena arising from the Earth's motion around the Sun, the Ptolemaic system included a secondary circle known as an epicycle, whose centre moved along the path of the primary orbital circle, known as the deferent. Ptolemy's Great Compilation, or Almagest after its Arabic translation, was to astronomy what Euclid's Elements was to geometry. Contrary to the Elements, however, the Almagest deploys geometry for the purpose of calculation. Among the items Ptolemy calculated was a table of chords, which correspond to the trigonometric sine function later introduced by Indian and Islamic mathematicians. The table of chords assisted the calculation of distances from angular measurements as a modern astronomer might do with the law of sines.

      The application of geometry to astronomy reframed the perennial Greek pursuit of the nature of truth. If a mathematical description fit the facts, as did Ptolemy's explanation of the unequal lengths of the seasons by the eccentricity of the Sun's orbit, should the description be taken as true of nature? The answer, with increasing emphasis, was “no.” Astronomers remarked that the eccentric orbit representing the Sun's annual motion could be replaced by a pair of circles, a deferent centred on the Earth and an epicycle the centre of which moved along the circumference of the deferent. That gave two observationally equivalent solar theories based on two quite different mechanisms. Geometry was too prolific of alternatives to disclose the true principles of nature. The Greeks, who had raised a sublime science from a pile of practical recipes, discovered that in reversing the process, in reapplying their mathematics to the world, they had no securer claims to truth than the Egyptian rope pullers.

Ancient geometry: cosmological and metaphysical
Pythagorean numbers and Platonic solids (Platonic solid)
 The Pythagoreans used geometrical figures to illustrate their slogan that all is number—thus their “triangular numbers” (n(n−1)/2), “square numbers” (n2), and “altar numbers” (n3), some of which are shown in the figure—>. This principle found a sophisticated application in Plato's creation story, the Timaeus, which presents the smallest particles, or “elements,” of matter as regular geometrical figures. Since the ancients recognized four or five elements at most, Plato sought a small set of uniquely defined geometrical objects to serve as elementary constituents. He found them in the only three-dimensional structures whose faces are equal regular polygons that meet one another at equal solid angles: the tetrahedron, or pyramid (with 4 triangular faces); the cube (with 6 square faces); the octahedron (with 8 equilateral triangular faces); the dodecahedron (with 12 pentagonal faces); and the icosahedron (with 20 equilateral triangular faces).

      The cosmology of the Timaeus had a consequence of the first importance for the development of mathematical astronomy. It guided Johannes Kepler (Kepler, Johannes) (1571–1630) to his discovery of the laws of planetary motion (Kepler's laws of planetary motion). Kepler deployed the five regular Platonic solids not as indicators of the nature and number of the elements but as a model of the structure of the heavens. In 1596 he published Prodromus Dissertationum Mathematicarum Continens Mysterium Cosmographicum (“Cosmographic Mystery”), in which each of the known six planets revolved around the Sun on spheres separated by the five Platonic solids. Although Tycho Brahe (Brahe, Tycho) (1546–1601), the world's greatest observational astronomer before the invention of the telescope, rejected the Copernican model of the solar system, he invited Kepler to assist him at his new observatory outside of Prague. In trying to resolve discrepancies between his original theory and Brahe's observations, Kepler made the capital discovery that the planets move in ellipses around the Sun as a focus.

Measuring the Earth and heavens
 Geometry offered Greek cosmologists not only a way to speculate about the structure of the universe but also the means to measure it. South of Alexandria and roughly on the same meridian of longitude is the village of Syene (modern Aswān), where the Sun stands directly overhead at noon on a midsummer day. At the same moment at Alexandria, the Sun's rays make an angle α with the tip of a vertical rod, as shown in the figure—>. Since the Sun's rays fall almost parallel on the Earth, the angle subtended by the arc l (representing the distance between Alexandria and Syene) at the centre of the Earth also equals α; thus the ratio of the Earth's circumference, C, to the distance, l, must equal the ratio of 360° to the angle α—in symbols, C:l = 360°:α. Eratosthenes (Eratosthenes of Cyrene) made the measurements, obtaining a value of about 5,000 stadia for l, which gave a value for the Earth's circumference of about 400,000 stadia. Because the accepted length of the Greek stadium varied locally, we cannot accurately determine Eratosthenes' margin of error. However, if we credit the ancient historian Plutarch's guess at Eratosthenes' unit of length, we obtain a value for the Earth's circumference of about 46,250 km—remarkably close to the modern value (about 15 percent too large), considering the difficulty in accurately measuring l and α. (See Sidebar: Measuring the Earth, Classical and Arabic.)

       Aristarchus Of Samos (c. 310–230 BC) has garnered the credit for extending the grip of number as far as the Sun. Using the Moon as a ruler and noting that the apparent sizes of the Sun and the Moon are about equal, he calculated values for his treatise “On the Sizes and Distances of the Sun and Moon.” The great difficulty of making the observations resulted in an underestimation of the solar distance about 20-fold—he obtained a solar distance, σ, roughly 1,200 times the Earth's radius, r. Possibly Aristarchus' inquiry into the relative sizes of the Sun, Moon, and Earth led him to propound the first heliocentric (“Sun-centred”) model of the universe.

      Aristarchus' value for the solar distance was confirmed by an astonishing coincidence. Ptolemy equated the maximum distance of the Moon in its eccentric orbit with the closest approach of Mercury riding on its epicycle; the farthest distance of Mercury with the closest of Venus; and the farthest of Venus with the closest of the Sun. Thus he could compute the solar distance in terms of the lunar distance and thence the terrestrial radius. His answer agreed with that of Aristarchus. The Ptolemaic conception of the order and machinery of the planets, the most powerful application of Greek geometry to the physical world, thus corroborated the result of direct measurement and established the dimensions of the cosmos for over a thousand years. As the ancient philosophers said, there is no truth in astronomy.

The post-classical period
Passage through Islam
 Two centuries after they broke out of their desert around Mecca, the followers of Muhammad occupied the lands from Persia to Spain and settled down to master the arts and sciences of the peoples they had conquered. They admired especially the works of the Greek mathematicians and physicians and the philosophy of Aristotle. By the late 9th century they were already able to add to the geometry of Euclid, Archimedes, and Apollonius. In the 10th century they went beyond Ptolemy. Stimulated by the problem of finding the effective orientation for prayer (the qiblah, or direction from the place of worship to Mecca), Islamic geometers and astronomers developed the stereographic projection (invented to project the celestial sphere onto a two-dimensional map or instrument) as well as plane and spherical trigonometry. Here they incorporated elements derived from India as well as from Greece. Their achievements in geometry and geometrical astronomy materialized in instruments for drawing conic sections and, above all, in the beautiful brass astrolabes (astrolabe) with which they reduced to the turn of a dial the toil of calculating astronomical quantities.

      Thābit ibn Qurrah (Thābit ibn Qurra) (836–901) had precisely the attributes required to bring the geometry of the Arabs up to the mark set by the Greeks. As a member of a religious sect close but hostile to both Jews and Christians, he knew Syriac and Greek as well as Arabic; as a money changer, he knew how to calculate; as both, he recommended himself to the Banū Mūsā, a set of mathematician brothers descended from a robber who had diversified into astrology. The Banū Mūsā directed a House of Wisdom in Baghdad sponsored by the caliph. There they presided over translations of the Greek classics. Thābit became an ornament of the House of Wisdom. He translated Archimedes and Apollonius, some of whose books now are known only in his versions. In a notable addition to Euclid, he tried valiantly to prove the parallel postulate (discussed later in Non-Euclidean geometries (geometry)).

      Among the pieces of Greek geometrical astronomy that the Arabs made their own was the planispheric astrolabe, which incorporated one of the methods of projecting the celestial sphere onto a two-dimensional surface invented in ancient Greece. One of the desirable mathematical features of this method (the stereographic projection) is that it converts circles into circles or straight lines, a property proved in the first pages of Apollonius's Conics. As Ptolemy showed in his Planisphaerium, the fact that the stereographic projection maps circles into circles or straight lines makes the astrolabe a very convenient instrument for reckoning time and representing the motions of celestial bodies. The earliest known Arabic astrolabes and manuals for their construction date from the 9th century. The Islamic world improved the astrolabe as an aid for determining the time for prayers, for finding the direction to Mecca, and for astrological divination.

Europe rediscovers the classics
      Contacts among Christians, Jews, and Arabs in Catalonia brought knowledge of the astrolabe to the West before the year 1000. During the 12th century many manuals for its use and construction were translated into Latin along with geometrical works by the Banū Mūsā, Thābit, and others. Some of the achievements of the Arab geometers were rediscovered in the West after wide and close study of Euclid's Elements, which was translated repeatedly from the Arabic and once from the Greek in the 12th and 13th centuries. The Elements (Venice, 1482) was one of the first technical books ever printed. Archimedes also came West in the 12th century, in Latin translations from Greek and Arabic sources. Apollonius arrived only by bits and pieces. Ptolemy's Almagest appeared in Latin manuscript in 1175. Not until the humanists of the Renaissance turned their classical learning to mathematics, however, did the Greeks come out in standard printed editions in both Latin and Greek.

      These texts affected their Latin readers with the strength of revelation. Europeans discovered the notion of proof, the power of generalization, and the superhuman cleverness of the Greeks; they hurried to master techniques that would enable them to improve their calendars and horoscopes, fashion better instruments, and raise Christian mathematicians to the level of the infidels. It took more than two centuries for the Europeans to make their unexpected heritage their own. By the 15th century, however, they were prepared to go beyond their sources. The most novel developments occurred where creativity was strongest, in the art of the Italian Renaissance.

      The theory of linear perspective, the brainchild of the Florentine architect-engineers Filippo Brunelleschi (Brunelleschi, Filippo) (1377–1446) and Leon Battista Alberti (Alberti, Leon Battista) (1404–72) and their followers, was to help remake geometry during the 17th century. The scheme of Brunelleschi and Alberti, as given without proofs in Alberti's De pictura (1435; On Painting), exploits the pyramid of rays that, according to what they had learned from the Westernized versions of the optics of Ibn al-Haytham (c. 965–1040), proceeds from the object to the painter's eye. Imagine, as Alberti directed, that the painter studies a scene through a window, using only one eye and not moving his head; he cannot know whether he looks at an external scene or at a glass painted to present to his eye the same visual pyramid. Supposing this decorated window to be the canvas, Alberti interpreted the painting-to-be as the projection of the scene in life onto a vertical plane cutting the visual pyramid. A distinctive feature of his system was the “point at infinity” at which parallel lines in the painting appear to converge.

      Alberti's procedure, as developed by Piero della Francesca (c. 1410–92) and Albrecht Dürer (Dürer, Albrecht) (1471–1528), was used by many artists who wished to render perspective persuasively. At the same time, cartographers tried various projections of the sphere to accommodate the record of geographical discoveries that began in the mid-15th century with Portuguese exploration of the west coast of Africa. Coincidentally with these explorations, mapmakers recovered Ptolemy's Geography, in which he had recorded by latitude (sometimes near enough) and longitude (usually far off) the principal places known to him and indicated how they could be projected onto a map.

      The discoveries that enlarged the known Earth did not fit easily on Ptolemy's projections. Cartographers therefore adopted the stereographic projection that had served astronomers. Several projected the Northern Hemisphere onto the Equator just as in the standard astrolabe, but the most widely used aspect, popularized in the world maps made by Gerardus Mercator (Mercator, Gerardus)'s son for later editions of his father's atlas (beginning in 1595), projected points on the Earth onto a cylinder tangent to the Earth at the Equator. After cutting the cylinder along a vertical line and flattening the resulting rectangle, the result was the now-familiar Mercator (Mercator projection) map.

      The intense cultivation of methods of projection by artists, architects, and cartographers during the Renaissance eventually provoked mathematicians into considering the properties of linear perspective in general. The most profound of these generalists was a sometime architect named Girard Desargues (Desargues, Girard) (1591–1661).

Transformation
French circles
      Desargues was a member of intersecting circles of 17th-century French mathematicians worthy of Plato's Academy of the 4th century BC or Baghdad's House of Wisdom of the 9th century AD. They included René Descartes (Descartes, René) (1596–1650) and Pierre de Fermat (Fermat, Pierre de) (1601–65), inventors of analytic geometry; Gilles Personne de Roberval (Roberval, Gilles Personne de) (1602–75), a pioneer in the development of the calculus; and Blaise Pascal (Pascal, Blaise) (1623–62), a contributor to the calculus and an exponent of the principles set forth by Desargues.

      Two main directions can be distinguished in Desargues's work. Like Renaissance artists, Desargues (Desargues, Girard) freely admitted the point at infinity into his demonstrations and showed that every set of parallel lines in a scene (apart from those parallel to the sides of the canvas) should project as converging bundles at some point on the “line at infinity” (the horizon). With the addition of points at infinity to the Euclidean plane, Desargues could frame all his propositions about straight lines without excepting parallel ones—which, like the others, now met one another, although not before “infinity.” A farther-reaching matter arising from artistic perspective was the relation between projections of the same object from different points of view and different positions of the canvas. Desargues observed that neither size nor shape is generally preserved in projections, but collinearity is, and he provided an example, possibly useful to artists, in images of triangles seen from different points of view. The statement that accompanied this example became known as Desargues's theorem.

 Desargues's second direction was to “simplify” Apollonius's work on conic sections (conic section). Despite his generality of approach, Apollonius needed to prove all his theorems for each type of conic separately. Desargues saw that he could prove them all at once and, moreover, by treating a cylinder as a cone with vertex at infinity, demonstrate useful analogies between cylinders and cones. Following his lead, Pascal made his surprising discovery that the intersections of the three pairs of opposite sides of a hexagon inscribed in a conic lie on a straight line. (See the figure—>.) In 1685, in his Sectiones Conicæ, Philippe de la Hire (1640–1718), a Parisian painter turned mathematician, proved several hundred propositions in Apollonius's Conics by Desargues's efficient methods.

Cartesian geometry
      In 1619, as part of the great illumination that inspired Descartes (Descartes, René) to assume the modest chore of reforming philosophy as well as mathematics, he devised “compasses” made of sticks sliding in grooved frames to duplicate the cube and trisect angles. Descartes esteemed these implements and the constructions they effected as (to quote from a letter of 1619) “no less certain and geometrical than the ordinary ones with which circles are drawn.” By the use of apt instruments, he would bring ancient mathematics to perfection: “scarcely anything will remain to be discovered in geometry.”

      What Descartes had in mind was the use of compasses with sliding members to generate curves. To classify and study such curves, Descartes took his lead from the relations Apollonius had used to classify conic sections, which contain the squares, but no higher powers, of the variables. To describe the more complicated curves produced by his instruments or defined as the loci of points satisfying involved criteria, Descartes had to include cubes and higher powers of the variables. He thus overcame what he called the deceptive character of the terms square, rectangle, and cube as used by the ancients and came to identify geometric curves as depictions of relationships defined algebraically. By reducing relations difficult to state and prove geometrically to algebraic relations between coordinates (usually rectangular) of points on curves, Descartes brought about the union of algebra and geometry that gave birth to the calculus.

Geometrical calculus
      The familiar use of infinity, which underlay much of perspective theory and projective geometry, also leavened the tedious Archimedean method of exhaustion. Not surprisingly, a practical man, the Flemish engineer Simon Stevin (Stevin, Simon) (1548–1620), who wrote on perspective and cartography among many other topics of applied mathematics, gave the first effective impulse toward redefining the object of Archimedean analysis. Instead of confining the circle between an inscribed and a circumscribed polygon, the new view regarded the circle as identical to the polygons, and the polygons to one another, when the number of their sides becomes infinitely great.

 This revitalized approach to exhaustion received a preliminary systematization in the Geometria Indivisibilibus Continuorum Nova Quadam Ratione Promota (1635; “A Method for the Determination of a New Geometry of Continuous Indivisibles”) by the Italian mathematician Bonaventura (Francesco) Cavalieri (Cavalieri, Bonaventura) (1598–1647). Cavalieri, perhaps influenced by Kepler's method of determining volumes in Nova Steriometria Doliorum (1615; “New Stereometry of Wine Barrels”), regarded lines as made up of an infinite number of dimensionless points, areas as made up of lines of infinitesimal thickness, and volumes as made up of planes of infinitesimal depth in order to obtain algebraic ways of summing the elements into which he divided his figures. Cavalieri's method may be stated as follows: if two figures (solids) of equal height are cut by parallel lines (planes) such that each pair of lengths (areas) matches, then the two figures (solids) have the same area (volume). (See the figure—>.) Although not up to the rigorous standards of today and criticized by “classicist” contemporaries (who were unaware that Archimedes himself had explored similar techniques), Cavalieri's method of indivisibles became a standard tool for solving volumes until the introduction of integral calculus near the end of the 17th century.

 A second geometrical inspiration for the calculus derived from efforts to define tangents to curves more complicated than conics. Fermat's (Fermat, Pierre de) method, representative of many, had as its exemplar the problem of finding the rectangle that maximizes the area for a given perimeter. Let the sides sought for the rectangle be denoted by a and b. Increase one side and diminish the other by a small amount ε; the resultant area is then given by (a + ε)(b − ε). Fermat observed what Kepler had perceived earlier in investigating the most useful shapes for wine casks, that near its maximum (or minimum) a quantity scarcely changes as the variables on which it depends alter slightly. On this principle, Fermat equated the areas ab and (a + ε)(b − ε) to obtain the stationary values: ab = ab − εa + εb − ε2. By canceling the common term ab, dividing by ε, and then setting ε at zero, Fermat had his well-known answer, a = b. The figure with maximum area is a square. To obtain the tangent to a curve by this method, Fermat began with a secant through two points a short distance apart and let the distance vanish (see the figure—>).

The world system
      Part of the motivation for the close study of Apollonius during the 17th century was the application of conic sections to astronomy. Kepler not only replaced the many circles of the old planetary system with a few ellipses, he also substituted a complicated rule of motion (his “second law”) for the relatively simple Ptolemaic rule that all motions must be compounded of rotations performed at constant velocity. Kepler's second law states that a planet moves in its ellipse so that the line between it and the Sun placed at a focus sweeps out equal areas in equal times. His astronomy thus made pressing and practical the otherwise merely difficult problem of the quadrature of conics and the associated theory of indivisibles.

      With the methods of Apollonius and a few infinitesimals, an inspired geometer showed that the laws regarding both area and ellipse can be derived from the suppositions that bodies free from all forces either rest or travel uniformly in straight lines and that each planet constantly falls toward the Sun with an acceleration that depends only on the distance between their centres. The inspired geometer was Isaac Newton (Newton, Sir Isaac) (1642 [Old Style]–1727), who made planetary dynamics a matter entirely of geometry by replacing the planetary orbit by a succession of infinitesimal chords, planetary acceleration by a series of centripetal jerks, and, in keeping with Kepler's second law, time by an area.

      Besides the problem of planetary motion, questions in optics pushed 17th-century natural philosophers and mathematicians to the study of conic sections. As Archimedes is supposed to have shown (or shone) in his destruction of a Roman fleet by reflected sunlight, a parabolic mirror brings all rays parallel to its axis to a common focus. The story of Archimedes provoked many later geometers, including Newton, to emulation. Eventually they created instruments powerful enough to melt iron.

      The figuring of telescope lenses likewise strengthened interest in conics after Galileo Galilei (Galileo)'s revolutionary improvements to the astronomical telescope in 1609. Descartes emphasized the desirability of lenses with hyperbolic surfaces, which focus bundles of parallel rays to a point (spherical lenses of wide apertures give a blurry image), and he invented a machine to cut them—which, however, proved more ingenious than useful.

      A final example of early modern applications of geometry to the physical world is the old problem of the size of the Earth. (See Sidebar: Measuring the Earth, Modernized.) On the hypothesis that the Earth cooled from a spinning liquid blob, Newton calculated that it is an oblate spheroid (obtained by rotating an ellipse around its minor axis), not a sphere, and he gave the excess of its equatorial over its polar diameter. During the 18th century many geodesists tried to find the eccentricity of the terrestrial ellipse. At first it appeared that all the measurements might be compatible with a Newtonian Earth. By the end of the century, however, geodesists had uncovered by geometry that the Earth does not, in fact, have a regular geometrical shape.

Relaxation and rigour
      The dominance of analysis (algebra and the calculus) during the 18th century produced a reaction in favour of geometry early in the 19th century. Fundamental new branches of the subject resulted that deepened, generalized, and violated principles of ancient geometry. The cultivators of these new fields, such as Jean-Victor Poncelet (Poncelet, Jean-Victor) (1788–1867) and his self-taught disciple Jakob Steiner (Steiner, Jakob) (1796–1863), vehemently urged the claims of geometry over analysis. The early 19th-century revival of pure geometry produced the discovery that Euclid had devoted his efforts to only one of several comprehensive geometries, the others of which can be created by replacing Euclid's fifth postulate with another about parallels.

Projection again
      Poncelet, who was an officer in the French corps of engineers, learned scraps of Desargues's work from his teacher Gaspard Monge (Monge, Gaspard, comte de Péluse) (1746–1818), who developed his own method of projection for drawings of buildings and machines. Poncelet relied on this information to keep himself alive. Taken captive during Napoleon's invasion of Russia in 1812, he passed his time by rehearsing in his head the things he had learned from Monge. The result was projective geometry.

      Poncelet employed three basic tools. One he took from Desargues: the demonstration of difficult theorems about a complicated figure by working out equivalent simpler theorems on an elementary figure interchangeable with the original figure by projection. The second tool, continuity, allows the geometer to claim certain things as true for one figure that are true of another equally general figure provided that the figures can be derived from one another by a certain process of continual change. Poncelet and his defender Michel Chasles (Chasles, Michel) (1793–1880) extended the principle of continuity into the domain of the imagination by considering constructs such as the common chord in two circles that do not intersect.

 Poncelet's third tool was the “principle of duality,” which interchanges various concepts such as points with lines, or lines with planes, so as to generate new theorems from old theorems. Desargues's theorem allows their interchange. So, as Steiner showed, does Pascal's theorem that the three points of intersection of the opposite sides of a hexagon inscribed in a conic lie on a line; thus, the lines joining the opposite vertices of a hexagon circumscribed about a conic meet in a point. (See the figure—>.)

      Poncelet's followers realized that they were hampering themselves, and disguising the true fundamentality of projective geometry, by retaining the concept of length and congruence in their formulations, since projections do not usually preserve them. Similarly, parallelism had to go. Efforts were well under way by the middle of the 19th century, by Karl George Christian von Staudt (Staudt, Karl Georg Christian von) (1798–1867) among others, to purge projective geometry of the last superfluous relics from its Euclidean past.

Non-Euclidean geometries
 The Enlightenment was not so preoccupied with analysis as to completely ignore the problem of Euclid's fifth postulate. In 1733 Girolamo Saccheri (1667–1733), a Jesuit professor of mathematics at the University of Pavia, Italy, substantially advanced the age-old discussion by setting forth the alternatives in great clarity and detail before declaring that he had “cleared Euclid of every defect” (Euclides ab Omni Naevo Vindicatus, 1733). Euclid's fifth postulate runs: “If a straight line falling on two straight lines makes the interior angles on the same side less than two right angles, the straight lines, if produced indefinitely, will meet on that side on which are the angles less than two right angles.” Saccheri took up the quadrilateral of Omar Khayyam (1048–1131), who started with two parallel lines AB and DC, formed the sides by drawing lines AD and BC perpendicular to AB, and then considered three hypotheses for the internal angles at C and D: to be right, obtuse, or acute (see the figure—>). The first possibility gives Euclidean geometry. Saccheri devoted himself to proving that the obtuse and the acute alternatives both end in contradictions, which would thereby eliminate the need for an explicit parallel postulate.

      On the way to this spurious demonstration, Saccheri established several theorems of non-Euclidean geometry—for example, that according to whether the right, obtuse, or acute hypothesis is true, the sum of the angles of a triangle respectively equals, exceeds, or falls short of 180°. He then destroyed the obtuse hypothesis by an argument that depended upon allowing lines to increase in length indefinitely. If this is disallowed, the hypothesis of the obtuse angle produces a system equivalent to standard spherical geometry, the geometry of figures drawn on the surface of a sphere.

      As for the acute angle, Saccheri could defeat it only by appealing to an arbitrary hypothesis about the behaviour of lines at infinity. One of his followers, the Swiss-German polymath Johann Heinrich Lambert (Lambert, Johann Heinrich) (1728–77), observed that, based on the acute hypothesis, the area of a triangle is the negative of that of a spherical triangle. Since the latter is proportional to the square of the radius, r, the former appeared to Lambert to be the area of an imaginary sphere with radius ir, where i = √(−1) .

      Although both Saccheri and Lambert aimed to establish the hypothesis of the right angle, their arguments seemed rather to indicate the unimpeachability of the alternatives. Several mathematicians at the University of Göttingen, notably the great Carl Friedrich Gauss (Gauss, Carl Friedrich) (1777–1855), then took up the problem. Gauss was probably the first to perceive that a consistent geometry could be built up independent of Euclid's fifth postulate, and he derived many relevant propositions, which, however, he promulgated only in his teaching and correspondence. The earliest published non-Euclidean geometric systems were the independent work of two young men from the East who had nothing to lose by their boldness. Both can be considered Gauss's disciples once removed: the Russian Nikolay Ivanovich Lobachevsky (Lobachevsky, Nikolay Ivanovich) (1792–1856), who learned his mathematics from a close friend of Gauss's at the University of Kazan, where Lobachevsky later became a professor; and János Bolyai (Bolyai, János) (1802–60), an officer in the Austro-Hungarian army whose father also was a friend of Gauss's. Both Lobachevsky and Bolyai had worked out their novel geometries by 1826.

      Lobachevsky and Bolyai reasoned about the hypothesis of the acute angle in the manner of Saccheri and Lambert and recovered their results about the areas of triangles. They advanced beyond Saccheri and Lambert by deriving an imaginary trigonometry to go with their imaginary geometry. Just as Desargues's projective geometry was neglected for many years, so the work of Bolyai and Lobachevsky made little impression on mathematicians for a generation and more. It was largely the posthumous publication in 1855 of Gauss's ideas about non-Euclidean geometry that gave the new approaches the cachet to attract the attention of later mathematicians.

A grand synthesis
      Another of the profound impulses Gauss gave geometry concerned the general description of surfaces. Typically—with the notable exception of the geometry of the sphere—mathematicians had treated surfaces as structures in three-dimensional Euclidean space. However, as these surfaces occupy only two dimensions, only two variables are needed to describe them. This prompted the thought that two-dimensional surfaces could be considered as “spaces” with their own geometries, not just as Euclidean structures in ordinary space. For example, the shortest distance, or path, between two points on the surface of a sphere is the lesser arc of the great circle joining them, whereas, considered as points in three-dimensional space, the shortest distance between them is an ordinary straight line.

      The shortest path between two points on a surface lying wholly within that surface is called a geodesic, which reflects the origin of the concept in geodesy, in which Gauss took an active interest. His initiative in the study of surfaces as spaces and geodesics as their “lines” was pursued by his student and, briefly, his successor at Göttingen, Bernhard Riemann (Riemann, Bernhard) (1826–66). Riemann began with an abstract space of n dimensions. That was in the 1850s, when mathematicians and mathematical physicists were beginning to use n-dimensional Euclidean space to describe the motions of systems of particles in the then-new kinetic theory of gases. Riemann worked in a quasi-Euclidean space—“quasi” because he used the calculus to generalize the Pythagorean theorem to supply sufficient flexibility to provide for geodesics on any surface.

 When this very general differential geometry came down to two-dimensional surfaces of constant curvature, it revealed excellent models for non-Euclidean geometries. Riemann himself pointed out that, merely by calling the geodesics of a sphere “straight lines,” the maligned hypothesis of the obtuse angle produces the geometry appropriate to the sphere's surface. Similarly, as shown by Eugenio Beltrami (Beltrami, Eugenio) (1835–1900), who ended his teaching career in Saccheri's old post at Pavia, the geometry defined in the plane by the hypothesis of the acute angle fits perfectly a surface of revolution of constant negative curvature now called a pseudosphere (see the figure—>)—again, provided that its geodesics are accepted as the straight lines of the geometry.

      Since the hypothesis of the obtuse angle correctly characterizes Euclidean geometry applied to the surface of a sphere, the non-Euclidean geometry based on it must be exactly as consistent as Euclidean geometry. The case of the acute angle treated by Lobachevsky and Bolyai required a sharper tool. Beltrami found it in a projection into a disc in the Euclidean plane of the points of a non-Euclidean space, in which each geodesic from the non-Euclidean space corresponds to a chord of the disc. Geometry built on the hypothesis of the acute angle has the same consistency as Euclidean geometry.

      The key role of Euclidean geometry in proofs of the consistency of non-Euclidean geometries exposed the Elements to ever-deeper scrutiny. The old blemishes—particularly appeals to intuition and diagrams for the meaning of concepts like “inside” and “between” and the use of questionable procedures like superposition to prove congruency—became intolerable to mathematicians who laboured to clarify the foundations of arithmetic and the calculus as well as the interrelations of the new geometries. The German mathematician Moritz Pasch (1843–1930), in his Vorlesungen über neuere Geometrie (1882; “Lectures on the New Geometry”), identified what was wanting: undefined concepts, axioms about those concepts, and more rigorous logic based on those axioms. The choice of undefined concepts and axioms is free, apart from the constraint of consistency. Mathematicians following Pasch's path introduced various elements and axioms and developed their geometries with greater or lesser elegance and trouble. The most successful of these systematizers was the Göttingen professor David Hilbert (Hilbert, David) (1862–1943), whose The Foundations of Geometry (1899) greatly influenced efforts to axiomatize all of mathematics. (See Elements.)

The real world
      Euclid's Elements had claimed the excellence of being a true account of space. Within this interpretation, Euclid's fifth postulate was an empirical finding; non-Euclidean geometries did not apply to the real world. Bolyai apparently could not free himself from the persuasion that Euclidean geometry represented reality. Lobachevsky observed that, if there were a star so distant that its parallax was not observable from the Earth's orbit, his geometry would be indistinguishable from Euclid's at the point where the parallax vanished. By his calculation, based on stellar parallaxes then just detected, his geometry could be physically meaningful only in gargantuan triangles spanning interstellar space.

      In fact, non-Euclidean geometries apply to the cosmos more locally than Lobachevsky imagined. In 1916 Albert Einstein (Einstein, Albert) (1879–1955) published “The Foundation of the General Theory of relativity,” which replaced Newton's description of gravitation as a force that attracts distant masses to each other through Euclidean space with a principle of least effort, or shortest (temporal) path, for motion along the geodesics of a curved space. Einstein not only explained how gravitating bodies give this surface its properties—that is, mass determines how the differential distances, or curvatures, in Riemann's geometry differ from those in Euclidean space—but also successfully predicted the deflection of light, which has no mass, in the vicinity of a star or other massive body. This was an extravagant piece of geometrizing—the replacement of gravitational force by the curvature of a surface. But it was not all. In relativity theory time is considered to be a dimension along with the three dimensions of space. On the closed four-dimensional world thus formed, the history of the universe stands revealed as describable by motion within a vast congeries of geodesics in a non-Euclidean universe.

J.L. Heilbron

Additional Reading

General history
The best overview in English of the history of geometry and its applications consists of the relevant chapters of Morris Kline, Mathematical Thought from Ancient to Modern Times (1972, reissued in 3 vol., 1990), which can be supplemented, for further applications, by Mathematics in Western Culture (1953, reissued 1987). Three other useful books of large scope are Petr Beckmann, A History of π, 4th ed. (1977, reissued 1993); Julian Lowell Coolidge, A History of Geometrical Methods (1940, reissued 1963); and David Wells, The Penguin Dictionary of Curious and Interesting Geometry (1991). A fine recent survey at a college level of the various branches of geometry, with much historical material, is David A. Brannan, Matthew F. Esplen, and Jeremy J. Gray, Geometry (1999).

Ancient Greek geometry
The standard English editions of the Greek geometers are those prepared by Thomas Little Heath beginning in the 1890s. They contain important historical and critical notes. Most exist in inexpensive reprints: Apollonius of Perga: Treatise on Conic Sections (1896, reissued 1961); The Works of Archimedes (1897, reissued 1953); Aristarchus of Samos, The Ancient Copernicus (1913, reprinted 1981); and The Thirteen Books of Euclid's Elements, 2nd ed., rev. with additions, 3 vol. (1926, reissued 1956). The historical material has been shortened and simplified, and its coverage extended, in A History of Greek Mathematics, 2 vol. (1921, reprinted 1993).Further information about technical-historical points—for example, the lunules of Hippocrates—may be found in Wilbur Richard Knorr, The Ancient Tradition of Geometric Problems (1986, reissued 1993). The epistemology of Greek geometry can be approached via the editor's introduction to and the text of Proclus, A Commentary on the First Book of Euclid's Elements, trans. and ed. by Glenn R. Morrow (1970, reprinted 1992).

Ancient non-Greek geometry
Other ancient geometrical traditions are covered in A.K. Bag, Mathematics in Ancient and Medieval India (1979); Richard J. Gillings, Mathematics in the Time of the Pharaohs (1972, reprinted 1982); Joseph Needham, Mathematics and the Sciences of the Heavens and the Earth (1959), vol. 3 of Science and Civilization in China; and B.L. van der Waerden, Science Awakening, 4th ed., 2 vol. (1975).

Geometry in Islam
Aspects of the extensive development of geometry by Islamic mathematicians can be studied in J.L. Berggren, Episodes in the Mathematics of Medieval Islam (1986). Otherwise, the best route to a survey is through the relevant chapters in vol. 2 of Roshdi Roshed (Rushdi Rashid) (ed.), Histoire des Sciences Arabes, 3 vol. (1997), and the articles on Arab mathematicians and astronomers in Charles Coulston Gillispie (ed.), Dictionary of Scientific Biography, 18 vol. (1970–90).

Renaissance geometry and applications
J.L. Heilbron, Geometry Civilized: History, Culture, and Technique (1998, reissued 2000), considers examples of geometry from some modern cultures as well as from the ancient Mediterranean and gives examples of the development of Greek geometry in the Middle Ages and Renaissance. A more advanced book along similar lines, but with more restricted coverage, is Alistair Macintosh Wilson, The Infinite in the Finite (1995). James Evans, The History and Practice of Ancient Astronomy (1998), is by far the best introduction to the theoretical and instrumental methods of the old astronomers. Albert van Helden, Measuring the Universe (1985), describes the methods of the Greeks and their development to the time of Halley. John P. Snyder, Flattening the Earth: Two Thousand Years of Map Projections (1993, reissued 1997), gives the neophyte cartographer a start. J.V. Field, The Invention of Infinity: Mathematics and Art in the Renaissance (1997), contains an elegant account, in both words and pictures, of the theory of projection of Brunelleschi, Alberti, and their followers.

Geometry and the calculus
The transformation of mathematics in the 17th century can be followed in Carl B. Boyer, The Concepts of the Calculus: A Critical and Historical Discussion (1939, reissued 1949; also published as The History of the Calculus and Its Conceptual Development, 1949, reissued 1959), largely superseded by Margaret E. Baron, The Origins of the Infinitesimal Calculus (1969, reprinted 1987); Michael S. Mahoney, The Mathematical Career of Pierre de Fermat (1601–65) (1973); and René Descartes, Discourse on Method, Optics, Geometry, and Meteorology, trans. by Paul J. Olscamp (1965, reissued 1976). This last work, which ranks among the most important books on natural philosophy and mathematics ever written, repays the effort required to master its idiom.

Axiomatic Euclidean and non-Euclidean geometry
Roberto Bonola, Non-Euclidean Geometry, 2nd rev. ed. (1938, reissued 1955), contains a thorough discussion of the work of Saccheri, Gauss, Bolyai, and Lobachevsky as well as a major text from each of the two founders of non-Euclidean geometry. David Hilbert, Foundations of Geometry, 2nd ed., trans. by Leo Unger and rev. and enlarged by Paul Bernays (1971, reissued 1992), is an excellent and accessible English translation.J.L. Heilbron

* * *


Universalium. 2010.

Игры ⚽ Нужна курсовая?

Look at other dictionaries:

  • Geometry — (Greek γεωμετρία ; geo = earth, metria = measure) is a part of mathematics concerned with questions of size, shape, and relative position of figures and with properties of space. Geometry is one of the oldest sciences. Initially a body of… …   Wikipedia

  • Geometry — Ge*om e*try, n.; pl. {Geometries}[F. g[ e]om[ e]trie, L. geometria, fr. Gr. ?, fr. ? to measure land; ge a, gh^, the earth + ? to measure. So called because one of its earliest and most important applications was to the measurement of the earth s …   The Collaborative International Dictionary of English

  • geometry — [jē äm′ə trē] n. pl. geometries [ME geometrie < OFr < L geometria < Gr geōmetria < geōmetrein, to measure the earth < gē, earth + metria, measurement < metrein, to measure: for IE base, see METER1] 1. the branch of mathematics… …   English World dictionary

  • geometry — early 14c., from O.Fr. géométrie (12c.), from L. geometria, from Gk. geometria measurement of earth or land; geometry, from comb. form of ge earth, land + metria (see METRY (Cf. metry)) …   Etymology dictionary

  • geometry — ► NOUN (pl. geometries) 1) the branch of mathematics concerned with the properties and relations of points, lines, surfaces, and solids. 2) the shape and relative arrangement of the parts of something. DERIVATIVES geometrician noun. ORIGIN Greek …   English terms dictionary

  • geometry — Although various laws concerning lines and angles were known to the Egyptians and the Pythagoreans, the systematic treatment of geometry by the axiomatic method began with the Elements of Euclid . From a small number of explicit axioms,… …   Philosophy dictionary

  • geometry — [[t]ʤiɒ̱mɪtri[/t]] 1) N UNCOUNT Geometry is the branch of mathematics concerned with the properties and relationships of lines, angles, curves, and shapes. ...the very ordered way in which mathematics and geometry describe nature. 2) N UNCOUNT:… …   English dictionary

  • Geometry — Mathematics Math e*mat ics, n. [F. math[ e]matiques, pl., L. mathematica, sing., Gr. ? (sc. ?) science. See {Mathematic}, and { ics}.] That science, or class of sciences, which treats of the exact relations existing between quantities or… …   The Collaborative International Dictionary of English

  • geometry — noun a) The branch of mathematics dealing with spatial relationships. spherical geometry b) A type of geometry with particular properties. See Also: geometer, geometrical …   Wiktionary

  • geometry — n. descriptive; Euclidean; plane; projective; solid geometry * * * [dʒɪ ɒmɪtrɪ] Euclidean plane projective solid geometry descriptive …   Combinatory dictionary

Share the article and excerpts

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