You are on page 1of 157

Aryabhatta the Indian mathematician

Aryabhatta, also known as Aryabhatta I or Aryabhata (476-550?), is a famous Indian


mathematician and astronomer, born in a place called Taregana, in Bihar (though some people do
not agree with the evidence). Taregana which literally means songs of stars in Bihari, is a small
place situated nearly 30 km from Patna, which was then known as Kusumpura later Pataliputra,
the capital of the Gupta Empire. This is the very empire that has been dubbed as the golden
period in Indian history. The best introduction to the genius of past is seen in the words of
Bhaskara I who said, Aryabhatta is the master who, after reaching the furthest shores and
plumbing the inmost depths of the sea of ultimate knowledge of mathematics, kinematics and
spherics, handed over the three sciences to the learned world.

Aryabhatta, the Indian mathematician

Influence of Aryabhatta on science and mathematics


Aryabhatta is considered to be one of the mathematicians who changed the course of
mathematics and astronomy to a great extent. He is known to have considerable influence on
Arabic science world too, where he is referred to as Arjehir. His notable contributions to the
world of science and mathematics includes the theory that the earth rotates on its axis,
explanations of the solar and lunar eclipses, solving of quadratic equations, place value system
with zero, and approximation of pie ().

Aryabhatta exerted influence on the Indian astronomical tradition to such an extent that his
presence was felt in neighboring countries and cultures also. There have been various
translations of his work among which the Arabic translation during the 820CE is very significant.

When mathematical students are confused with trigonometry even today, Aryabhatta had defined
sine, cosine, versine and inverse sine back in his era, influencing the birth of trigonometry. The
signs were originally known as jya, kojya, utkrama-jya and otkram jya. In Arabic they were
translated as jiba and kojiba, which later when being translated into Latin was misunderstood to
be fold in a garment by Gerard of Cremona, who stated it as sinus, which meant fold in Latin.
Aryabhatta was the first mathematician to detail both sine and versine (1 cos x) tables, in 3.75
intervals from 0 to 90, to 4 decimal places.

Aryabhattas astronomical calculations influenced the Arabians, who used the trigonometric
tables to compute many astronomical tables. His calendared calculation has been in continuous
use in India, on which the present day Panchangam is based. His studies are also base for the
national calendars of Iran and Afghanistan today.

Aryabhatta Worlds greatest inventions

Aryabhatiya
It is known that Aryabhatta has authored at least three astronomical books, in addition he also
wrote some free stanzas. Among them Aryabhatiya is the only text that has survived to this
day, whereas unfortunately his other works have been extinct. It is a small treatise written is 118
verses, which summarizes the Hindu mathematics of that time. This great mathematical
masterpiece of the past starts with 10 verse introduction, which is then followed by mathematical
section which is written in 33 verses that gives out 66 mathematical rules, but there is no proof to
go with it. The mathematical part of the Aryabhatiya is about algebra, arithmetic, plane
trigonometry and spherical trigonometry in addition to advanced mathematics on continued
fractions, quadratic equations, sums of power series and a table of sines.

The next section consists of 25 verses which gives us glimpse into the planetary models. The
final section of the book is dedicated to sphere and eclipses which runs into 50 verses. He states
that the moon and planets shine by reflected sunlight. Instead of the prevailing cosmogony where
eclipses were believed to be caused by pseudo-planetary nodes Rahu and Ketu, he explains
eclipses in terms of shadows cast by earth or those shadows that fall on earth. It is amazing how
Aryabhatta could explain both lunar and solar eclipse so accurately.
Statue of Aryabhatta at Inter-University Centre for Astronomy and Astrophysics at Pune (India)

There is some argument over the claim of Aryabhatta being the inventor of place value system
that made use of zero. Georges Ifrah, in his work Universal history of numbers: From prehistory
to the invention of the computer (London, 1998) writes in work, ..it is extremely likely that
Aryabhatta knew the sign for zero and the numerals of the place value system. Georges Ifrah
has studied the works of Aryabhatta and found that the counting and mathematical work carried
out by him would have been not possible without zero or place value system.

Honouring Aryabhatta
The Indian ISRO (Indian Space Research Organization) named its first satellite after the genius
mathematician and astronomer. A research establishment has been set up in Nainital, called the
Aryabhatta Research Institute of Observational Sciences (ARIOS) to honor his contribution to
the field of science. There is also a lunar crater and a species of bacteria discovered by ISRO
named after Aryabhatta.

Some of the works of Aryabhatta include


Aryabhatta worked out the value of pi.

He worked out the area of a triangle. His exact words were, ribhujasya phalashariram
samadalakoti bhujardhasamvargah which translates for a triangle, the result of a
perpendicular with the half side is the area.

He discussed the idea of sin.


He worked on the summation of series of squares and cubes (square-root and cube-root).

He talks about the rule of three which is to find the value of x when three numbers a, b
and c is given.

Aryabhatta calculates the volume of a sphere.

Aryabhatta described the model of the solar system, where the sun and moon are each
carried by epicycles that in turn revolve around the Earth. He also talks about the number
of rotations of the earth, describes that the earth rotating on its axis, the order of the
planets in terms of distance from earth.

Aryabhatta describes the solar and lunar eclipses were scientifically.

Aryabhatta describes that the moon and planets shine by light reflected from the sun.

Aryabhatta calculated the sidereal rotation which is the rotation of the earth with respect
to the stars as 23 hours, 56 minutes and 4.1 seconds.

He calculated the length of the sidereal year as 365 days, 6 hours, 12 minutes and 30
seconds. The actually value shows that his calculations was an error of 3 minutes and 20
seconds over a year.

Although we know nothing about the personal history of Aryabhatta, he was the genius who
continues to baffle mathematicians even to this day.

Thales of Miletus

Little is known of Thales. He was born about 624 BC in Miletus, Asia Minor (now Turkey) and
died about 546 BC in Miletos, Turkey
The bust shown above is in the Capitoline Museum in Rome but is not contemporary with
Thales.

Thales of Miletus

Some impression and highlights of his life and work follow:

Thales of Miletus was the first known Greek philosopher, scientist and mathematician.
Some consider him to be the teacher of of Pythagoras, though it may be only that he
advised Pythagoras to travel to Egypt and Chaldea.

From Eudemus of Rhodes (fl ca. 320 B.C) we know that he studied in Egypt and brought
these teachings to Greece. He is unanimously ascribed the introduction of mathematical
and astronomical sciences into Greece.

He is unanimously regarded as having been unusally clever--by general agreement the


first of the Seven Wise Men, a pupil of the Egyptians and the Chaldeans.

None of his writing survives; this makes it is difficult to determine his philosophy and to
be certain about his mathematical discoveries.

There is, of course, the story of his successful speculation in oil presses -- as testament to
his practical business acumen.

It is reported that he predicted an eclipse of the Sun on May 28, 585 BC, startling all of
Ionia.

He is credited with five theorems of elementary geometry.


From W. K. C. Guthrie we have

The achievement of Thales, has been represented by historians in two entirely different lights: on
the one hand, as a marvelous anticipation of modern scientific thinking, and on the other as
nothing but a transparent rationalization of a myth.

According to Guthrie himself, one may say that ``ideas of Thales and other Milesians created a
bridge between the two worlds-the world of myth and the world of the mind."

Thales believed that the Earth is a flat disk that floats on an endless expanse of water and all
things come to be from water.

But, more preciesly, Thales and the Milesians proceeded from the assumption of a fundamental
unity of all material things that is to be found behind their apparent diversity. This is the first
recorded monism in history. He also regards the world as alive and thus life and matter to be
inseparable. Even plants he feels have a immortal ``soul".

Being asked what was very difficult, he answered, in a famous apophthegm, "To Know
Thyself." Asked what was very easy, he answered, "To give advice." To the question, what/who
is God?, he answered, "That which has no beginning or no end." (The infinite!!)

So the task of the philosophers was to establish what exactly provided this unity: one said it was
water; another, the Boundless; yet another, air.

Thales is believed to have been the teacher of Anaximander and he is the first natural
philosopher in the Ionian (Milesian) School.

Thales is also said to have discovered a method of measuring the distance to a ship at sea.
Five basic propositions with proofs of plane geometry are attributed to Thales.

Proposition. A circle is bisected by any diameter.


Proposition. The base angles of an isosceles triangle are equal.
Proposition. The angles between two intersecting straight lines are equal.
Proposition. Two triangles are congruent if they have two angles and the included side
equal.
Proposition. An angle in a semicircle is a right angle.

Thales the Mathematician

Proposition. An angle in a semicircle is a right angle.

Proof.

Since there was no clear theory of angles at that time this is no doubt not the proof furnished by
Thales.
This article is about the Greek philosopher. For the mythological characters called Hippasus, see
Hippasus (mythology).

Hippasus of Metapontum

Hippasus (Ancient Greek: , Hppasos; 5th century BC) of Metapontum in Magna


Graecia, was a Pythagorean philosopher. Little is known about his life or his beliefs, but he is
sometimes credited with the discovery of the existence of irrational numbers. The discovery of
irrational numbers is said to have been shocking to the Pythagoreans, and Hippasus is supposed
to have drowned at sea, apparently as a punishment from the gods, for divulging this. However,
the few ancient sources which describe this story either do not mention Hippasus by name or
alternatively tell us that Hippasus drowned because he revealed how to construct a dodecahedron
inside a sphere. The discovery of irrationality is not specifically ascribed to Hippasus by any
ancient writer. Some modern scholars though have suggested that he discovered the irrationality
of 2, which it is believed was discovered around the time that he lived. But some sources
suggest that he was the one responsible for the revelation of the secret society of Pythagoreans.
Contents
1 Life

2 Doctrines

o 2.1 Irrational numbers

o 2.2 Proof of the irrationality of 2

3 References

4 See also

5 External links

Life
Little is known about the life of Hippasus. He may have lived in the late 5th century BC, about a
century after the time of Pythagoras. He probably came from Metapontum in Italy (Magna
Graecia), although the nearby city of Croton is also mentioned as his birthplace.[1] Iamblichus
states that he was the founder of a sect of the Pythagoreans called the Mathematici (Greek:
) in opposition to the Acusmatici (Greek: );[2] but elsewhere he makes
him the founder of the Acusmatici in opposition to the Mathematici.[3]

Doctrines
Aristotle speaks of Hippasus as holding the element of fire to be the cause of all things;[4] and
Sextus Empiricus contrasts him with the Pythagoreans in this respect, that he believed the arche
to be material, whereas they thought it was incorporeal, namely, number.[5] Diogenes Lartius
tells us that Hippasus believed that "there is a definite time which the changes in the universe
take to complete, and that the universe is limited and ever in motion."[6] According to one
statement, Hippasus left no writings,[7] according to another he was the author of the Mystic
Discourse, written to bring Pythagoras into disrepute.[8]

A scholium on Plato's Phaedo notes him as an early experimenter in music theory, claiming that
he made use of bronze disks to discover the fundamental musical ratios, 4:3, 3:2, and 2:1.[9]

Irrational numbers

Hippasus is sometimes credited with the discovery of the existence of irrational numbers,
following which he was drowned at sea. Pythagoreans preached that all numbers could be
expressed as the ratio of integers, and the discovery of irrational numbers is said to have shocked
them. However, the evidence linking the discovery to Hippasus is confused.

Pappus merely says that the knowledge of irrational numbers originated in the Pythagorean
school, and that the member who first divulged the secret perished by drowning.[10] Iamblichus
gives a series of inconsistent reports. In one story he explains how a Pythagorean was merely
expelled for divulging the nature of the irrational; but he then cites the legend of the Pythagorean
who drowned at sea for making known the construction of the regular dodecahedron in the
sphere.[11] In another account he tells how it was Hippasus who drowned at sea for betraying the
construction of the dodecahedron and taking credit for this construction himself;[12] but in another
story this same punishment is meted out to the Pythagorean who divulged knowledge of the
irrational.[13] Iamblichus clearly states that the drowning at sea was a punishment from the gods
for impious behaviour.[11]

These stories are usually taken together to ascribe the discovery of irrationals to Hippasus, but
whether he did or not is uncertain.[14] In principle, the stories can be combined, since it is possible
to discover irrational numbers when constructing dodecahedrons. Irrationality, by infinite
reciprocal subtraction, can be easily seen in the Golden ratio of the regular pentagon.[15]

Some modern scholars prefer to credit Hippasus with the discovery of the irrationality of 2.
Plato in his Theaetetus,[16] describes how Theodorus of Cyrene (c. 400 BC) proved the
irrationality of 3, 5, etc. up to 17, which implies that an earlier mathematician had already
proved the irrationality of 2.[17] A simple proof of the irrationality of 2 is indicated by Aristotle,
and it is set out in the proposition interpolated at the end of Euclid's Book X,[18] which suggests
that the proof was certainly ancient.[19] The proof is one of reductio ad absurdum, and the method
is to show that, if the diagonal of a square is commensurable with the side, then the same number
must be both odd and even.[19]

In the hands of modern writers this combination of vague ancient reports and modern guesswork
has sometimes evolved into a much more emphatic and colourful tale. Some writers have
Hippasus making his discovery while on board a ship, as a result of which his Pythagorean
shipmates toss him overboard;[20] while one writer even has Pythagoras himself "to his eternal
shame" sentencing Hippasus to death by drowning, for showing "that 2 is an irrational
number."[21]

Proof of the irrationality of 2

The reductio ad absurdum proof of the irrationality of 2 is as follows:[22]

1. Take a right triangle whose short sides are 1 unit in length

2. By the Pythagorean theorem, the diagonal is 2

3. Suppose that 2 is the ratio of two natural numbers, 2=m/n

4. Suppose that m/n has been reduced to its lowest common form by division
5. It follows that either m and n are both odd, or that m is odd and n is even, or that m is
even and n is odd (if not, we could reduce m/n even further by dividing both numbers by
2)

6. Square both sides of 2=m/n, so that 2=m2/n2

7. Then 2n2=m2, so that m2 is even, and therefore m is even

8. If m is even, then m=2x, where x is some other natural number

9. Squaring this, it follows that m2=4x2=2n2

10. It follows that n2=2x2, and therefore n2 is even, which means that n, being a natural
number, must be even

11. So we've reached a contradiction: although we assumed that m and n cannot both be
even, it now turns out they both are. It therefore follows that 2 cannot be expressed as
the ratio of two natural numbers, and must therefore be in another class of numbers

Anaxagoras (Ancient Greek: , Anaxagoras, "lord of the assembly"; c. 500


428 BC) was a Pre-Socratic Greek philosopher. Born in Clazomenae in Asia Minor, Anaxagoras
was the first philosopher to bring philosophy from Ionia to Athens. He attempted to give a
scientific account of eclipses, meteors, rainbows, and the sun, which he described as a fiery mass
larger than the Peloponnese. According to Diogenes Laertius and Plutarch, he fled to Lampsacus
due to a backlash against his pupil Pericles.

Anaxagoras is famous for introducing the cosmological concept of Nous (mind), as an ordering
force. He regarded material substance as an infinite multitude of imperishable primary elements,
referring all generation and disappearance to mixture and separation, respectively.

Contents
1 Biography

2 Cosmological theory

3 Literary references

4 See also

5 Notes

6 References

7 Further reading

8 External links

Biography
Anaxagoras appears to have had some amount of property and prospects of political influence in
his native town of Clazomenae in Asia Minor. However, he supposedly surrendered both of these
out of a fear that they would hinder his search for knowledge. Valerius Maximus preserves a
different tradition: Anaxagoras, coming home from a long voyage, found his property in ruin,
and said: "If this had not perished, I would have." A sentence, denoted by Maximus, as being
"possessed of sought-after wisdom!"[1][2] Although a Greek, he may have been a soldier of the
Persian army when Clazomenae was suppressed during the Ionian Revolt.

In early manhood (c. 464461 BC) he went to Athens, which was rapidly becoming the centre of
Greek culture. There he is said to have remained for thirty years. Pericles learned to love and
admire him, and the poet Euripides derived from him an enthusiasm for science and humanity.

Anaxagoras brought philosophy and the spirit of scientific inquiry from Ionia to Athens. His
observations of the celestial bodies and the fall of meteorites led him to form new theories of the
universal order. He attempted to give a scientific account of eclipses, meteors, rainbows, and the
sun, which he described as a mass of blazing metal, larger than the Peloponnese. He was the first
to explain that the moon shines due to reflected light from the sun. He also said that the moon
had mountains and believed that it was inhabited. The heavenly bodies, he asserted, were masses
of stone torn from the earth and ignited by rapid rotation. He explained that, though both sun and
the stars were fiery stones, we do not feel the heat of the stars because of their enormous distance
from earth. He thought that the earth is flat and floats supported by 'strong' air under it and
disturbances in this air sometimes causes earthquakes.[3] These speculations made him vulnerable
in Athens to a charge of impiety. Diogenes Laertius reports the story that he was prosecuted by
Cleon for impiety, but Plutarch says that Pericles sent his former tutor, Anaxagoras, to
Lampsacus for his own safety after the Athenians began to blame him for the Peloponnesian war.
[4]

About 450 BC, according to Laertius, Pericles spoke in defense of Anaxagoras at his trial.[5] Even
so, Anaxagoras was forced to retire from Athens to Lampsacus in Troad (c. 434433 BC). He
died there in around the year 428 BC. Citizens of Lampsacus erected an altar to Mind and Truth
in his memory, and observed the anniversary of his death for many years.

Anaxagoras wrote a book of philosophy, but only fragments of the first part of this have
survived, through preservation in work of Simplicius of Cilicia in the sixth century AD.

Cosmological theory

Anaxagoras, depicted as a medieval scholar in the Nuremberg Chronicle

All things have existed from the beginning. But originally they existed in infinitesimally small
fragments of themselves, endless in number and inextricably combined. All things existed in this
mass, but in a confused and indistinguishable form. There were the seeds (spermata) or
miniatures of wheat and flesh and gold in the primitive mixture; but these parts, of like nature
with their wholes (the homoiomereiai of Aristotle), had to be eliminated from the complex mass
before they could receive a definite name and character.

Mind arranged the segregation of like from unlike; panta chremata en omou eita nous elthon
auta diekosmese. This peculiar thing, called Mind (Nous), was no less illimitable than the chaotic
mass, but, unlike the logos of Heraclitus, it stood pure and independent (mounos ef eoutou), a
thing of finer texture, alike in all its manifestations and everywhere the same. This subtle agent,
possessed of all knowledge and power, is especially seen ruling in all the forms of life.[citation needed]

Mind causes motion. It rotated the primitive mixture, starting in one corner or point, and
gradually extended until it gave distinctness and reality to the aggregates of like parts, working
something like a centrifuge, and eventually creating the known cosmos. But even after it had
done its best, the original intermixture of things was not wholly overcome. No one thing in the
world is ever abruptly separated, as by the blow of an axe, from the rest of things.
It is noteworthy that Socrates (Plato, Phaedo, 98 B) accuses Anaxagoras of failing to
differentiate between nous and psyche, while Aristotle (Metaphysics, Book I) objects that his
nous is merely a deus ex machina to which he refuses to attribute design and knowledge.

Anaxagoras proceeded to give some account of the stages in the process from original chaos to
present arrangements. The division into cold mist and warm ether first broke the spell of
confusion. With increasing cold, the former gave rise to water, earth and stones. The seeds of life
which continued floating in the air were carried down with the rains and produced vegetation.
Animals, including man, sprang from the warm and moist clay. If these things be so, then the
evidence of the senses must be held in slight esteem. We seem to see things coming into being
and passing from it; but reflection tells us that decease and growth only mean a new aggregation
(synkrisis) and disruption (diakrisis). Thus, Anaxagoras distrusted the senses, and gave the
preference to the conclusions of reflection. Thus, he maintained that there must be blackness as
well as whiteness in snow; how, otherwise, could it be turned into dark water?

Anaxagoras marked a turning-point in the history of philosophy. With him, speculation passes
from the colonies of Greece to settle at Athens. By the theory of minute constituents of things,
and his emphasis on mechanical processes in the formation of order, he paved the way for the
atomic theory.

Literary references
Anaxagoras appears as a character in The Ionia Sanction, by Gary Corby

Squaring the circle is a problem proposed by ancient geometers. It is the challenge of


constructing a square with the same area as a given circle by using only a finite number of steps
with compass and straightedge. More abstractly and more precisely, it may be taken to ask
whether specified axioms of Euclidean geometry concerning the existence of lines and circles
entail the existence of such a square.
In 1882, the task was proven to be impossible, as a consequence of the LindemannWeierstrass
theorem which proves that pi () is a transcendental, rather than an algebraic irrational number;
that is, it is not the root of any polynomial with rational coefficients. It had been known for some
decades before then that the construction would be impossible if pi were transcendental, but pi
was not proven transcendental until 1882. Approximate squaring to any given non-perfect
accuracy, in contrast, is possible in a finite number of steps, since there are rational numbers
arbitrarily close to .

The expression "squaring the circle" is sometimes used as a metaphor for trying to do the
impossible.[1]

The term quadrature of the circle is sometimes used synonymously, or may refer to approximate
or numerical methods for finding the area of a circle.

Contents
1 History

2 Impossibility

3 Modern approximative constructions

4 Squaring or quadrature as integration

5 "Squaring the circle" as a metaphor

6 Claims of circle squaring

o 6.1 Connection with the longitude problem

o 6.2 Other modern claims

7 In literature

o 7.1 Literature

8 See also

9 References

10 External links
History
Methods to approximate the area of a given circle with a square were known already to
Babylonian mathematicians. The Egyptian Rhind papyrus of 1800BC gives the area of a circle as
(64/81) d 2, where d is the diameter of the circle, and pi approximated to 256/81, a number that
appears in the older Moscow Mathematical Papyrus, and used for volume approximations (i.e.
hekat (volume unit)). Indian mathematicians also found an approximate method, though less
accurate, documented in the Sulba Sutras.[2] Archimedes showed that the value of pi lay between
3 + 1/7 (approximately 3.1429) and 3 + 10/71 (approximately 3.1408). See Numerical
approximations of for more on the history.

The first Greek to be associated with the problem was Anaxagoras, who worked on it while in
prison. Hippocrates of Chios squared certain lunes, in the hope that it would lead to a solution
see Lune of Hippocrates. Antiphon the Sophist believed that inscribing regular polygons within a
circle and doubling the number of sides will eventually fill up the area of the circle, and since a
polygon can be squared, it means the circle can be squared. Even then there were skeptics
Eudemus argued that magnitudes cannot be divided up without limit, so the area of the circle will
never be used up.[3] The problem was even mentioned in Aristophanes's play The Birds.

It is believed that Oenopides was the first Greek who required a plane solution (that is, using
only a compass and straightedge). James Gregory attempted a proof of its impossibility in Vera
Circuli et Hyperbolae Quadratura (The True Squaring of the Circle and of the Hyperbola) in
1667. Although his proof was incorrect, it was the first paper to attempt to solve the problem
using algebraic properties of pi. It was not until 1882 that Ferdinand von Lindemann rigorously
proved its impossibility.

A partial history by Florian Cajori of attempts at the problem.[4]


The famous Victorian-age mathematician, logician and author, Charles Lutwidge Dodgson
(better known under the pseudonym, "Lewis Carroll") also expressed interest in debunking
illogical circle-squaring theories. In one of his diary entries for 1855, Dodgson listed books he
hoped to write including one called "Plain Facts for Circle-Squarers". In the introduction to "A
New Theory of Parallels", Dodgson recounted an attempt to demonstrate logical errors to a
couple of circle-squarers, stating:[5]

"The first of these two misguided visionaries filled me with a great ambition to do a feat I have
never heard of as accomplished by man, namely to convince a circle squarer of his error! The
value my friend selected for Pi was 3.2: the enormous error tempted me with the idea that it
could be easily demonstrated to BE an error. More than a score of letters were interchanged
before I became sadly convinced that I had no chance."

Impossibility
The solution of the problem of squaring the circle by compass and straightedge demands
construction of the number , and the impossibility of this undertaking follows from the fact
that pi is a transcendental (non-algebraic and therefore non-constructible) number. If the problem
of the quadrature of the circle is solved using only compass and straightedge, then an algebraic
value of pi would be found, which is impossible. Johann Heinrich Lambert conjectured that pi
was transcendental in 1768 in the same paper he proved its irrationality, even before the
existence of transcendental numbers was proven. It was not until 1882 that Ferdinand von
Lindemann proved its transcendence.

The transcendence of pi implies the impossibility of exactly "circling" the square, as well as of
squaring the circle.

It is possible to construct a square with an area arbitrarily close to that of a given circle. If a
rational number is used as an approximation of pi, then squaring the circle becomes possible,
depending on the values chosen. However, this is only an approximation and does not meet the
constraints of the ancient rules for solving the problem. Several mathematicians have
demonstrated workable procedures based on a variety of approximations.

Bending the rules by allowing an infinite number of compass-and-straightedge operations or by


performing the operations on certain non-Euclidean spaces also makes squaring the circle
possible. For example, although the circle cannot be squared in Euclidean space, it can be in
GaussBolyaiLobachevsky space. Indeed, even the preceding phrase is overoptimistic.[6][7]
There are no squares as such in the hyperbolic plane, although there are regular quadrilaterals,
meaning quadrilaterals with all sides congruent and all angles congruent (but these angles are
strictly smaller than right angles). There exist, in the hyperbolic plane, (countably) infinitely
many pairs of constructible circles and constructible regular quadrilaterals of equal area.
However, there is no method for starting with a regular quadrilateral and constructing the circle
of equal area, and there is no method for starting with a circle and constructing a regular
quadrilateral of equal area (even when the circle has small enough radius such that a regular
quadrilateral of equal area exists).
Modern approximative constructions
Though squaring the circle is an impossible problem using only compass and straightedge,
approximations to squaring the circle can be given by constructing lengths close to pi. It takes
only minimal knowledge of elementary geometry to convert any given rational approximation of
pi into a corresponding compass-and-straightedge construction, but constructions made in this
way tend to be very long-winded in comparison to the accuracy they achieve. After the exact
problem was proven unsolvable, some mathematicians applied their ingenuity to finding elegant
approximations to squaring the circle, defined roughly and informally as constructions that are
particularly simple among other imaginable constructions that give similar precision.

Among the modern approximate constructions was one by E. W. Hobson in 1913.[8] This was a
fairly accurate construction which was based on constructing the approximate value of
3.14164079..., which is accurate to 4 decimals (i.e. it differs from pi by about 4.8105).

Indian mathematician Srinivasa Ramanujan in 1913, C. D. Olds in 1963, Martin Gardner in


1966, and Benjamin Bold in 1982 all gave geometric constructions for

which is accurate to six decimal places of pi.

Kochaski's approximate construction

Srinivasa Ramanujan in 1914 gave a ruler-and-compass construction which was equivalent to


taking the approximate value for pi to be

giving a remarkable eight decimal places of pi.

In 1991, Robert Dixon gave constructions for


(Kochaski's approximation), though these were only accurate to four decimal places of pi.

Squaring or quadrature as integration


The problem of finding the area under a curve, known as integration in calculus, or quadrature in
numerical analysis, was known as squaring before the invention of calculus. Since the
techniques of calculus were unknown, it was generally presumed that a squaring should be done
via geometric constructions, that is, by compass and straightedge. For example Newton wrote to
Oldenberg in 1676 "I believe M. Leibnitz will not dislike the Theorem towards the beginning of
my letter pag. 4 for squaring Curve lines Geometrically" (emphasis added).[9] After Newton and
Leibniz invented calculus, they still referred to this integration problem as squaring a curve.

"Squaring the circle" as a metaphor


The futility of exercises aimed at finding the quadrature of the circle has lent itself to metaphors
describing a hopeless, meaningless, or vain undertaking.

For example, in Spanish, the expression "descubriste la cuadratura del crculo" ("you discovered
the quadrature of the circle") is often used derisively to dismiss claims that someone has found a
simple solution to a particularly hard or intractable problem.[citation needed]

Claims of circle squaring


Connection with the longitude problem

The mathematical proof that the quadrature of the circle is impossible using only compass and
straightedge has not proved to be a hindrance to the many people who have invested years in this
problem anyway. Having squared the circle is a famous crank assertion. (See also
pseudomathematics.) In his old age, the English philosopher Thomas Hobbes convinced himself
that he had succeeded in squaring the circle.

During the 18th and 19th century, the notion that the problem of squaring the circle was
somehow related to the longitude problem seems to have become prevalent among would-be
circle squarers. Using "cyclometer" for circle-squarer, Augustus de Morgan wrote in 1872:

Montucla says, speaking of France, that he finds three notions prevalent among cyclometers: 1.
That there is a large reward offered for success; 2. That the longitude problem depends on that
success; 3. That the solution is the great end and object of geometry. The same three notions are
equally prevalent among the same class in England. No reward has ever been offered by the
government of either country.[10]
Although from 1714 to 1828 the British government did indeed sponsor a 20,000 prize for
finding a solution to the longitude problem, exactly why the connection was made to squaring
the circle is not clear; especially since two non-geometric methods (the astronomical method of
lunar distances and the mechanical chronometer) had been found by the late 1760s. De Morgan
goes on to say that "[t]he longitude problem in no way depends upon perfect solution; existing
approximations are sufficient to a point of accuracy far beyond what can be wanted." In his book,
de Morgan also mentions receiving many threatening letters from would-be circle squarers,
accusing him of trying to "cheat them out of their prize".

Other modern claims

Even after it had been proved impossible, in 1894, amateur mathematician Edwin J. Goodwin
claimed that he had developed a method to square the circle. The technique he developed did not
accurately square the circle, and provided an incorrect area of the circle which essentially
redefined pi as equal to 3.2. Goodwin then proposed the Indiana Pi Bill in the Indiana state
legislature allowing the state to use his method in education without paying royalties to him. The
bill passed with no objections in the state house, but the bill was tabled and never voted on in the
Senate, amid increasing ridicule from the press.

In literature
Literature

The problem of squaring the circle has been mentioned by poets such as Dante and Alexander
Pope.

The character Meton of Athens in the play The Birds by Aristophanes (first performed in 414
BCE) mentions squaring the circle.

Dante's Paradise canto XXXIII lines 133-135 contain the verses:

As the geometer his mind applies


To square the circle, nor for all his wit
Finds the right formula, howe'er he tries

Pope, in his 1743 poem Dunciad, wrote:

Mad Mathesis alone was unconfined,


Too mad for mere material chains to bind,
Now to pure space lifts her ecstatic stare,
Now, running round the circle, finds it square.
Angle trisection
From Wikipedia, the free encyclopedia
(Redirected from Trisecting the angle)
Jump to: navigation, search

Angles may be trisected via a Neusis construction, but this uses tools outside the Greek
framework of an unmarked straightedge and a compass.

Angle trisection is a classic problem of compass and straightedge constructions of ancient Greek
mathematics. It concerns construction of an angle equal to one-third of a given arbitrary angle,
using only two tools: an un-marked straightedge, and a compass.

The problem as stated is generally impossible to solve, as shown by Pierre Wantzel (1837).
Wantzel's proof relies on ideas from the field of Galois theoryin particular, trisection of an
angle corresponds to the solution of a certain cubic equation, which is not possible using the
given tools. Note that the fact that there is no way to trisect an angle in general with just a
compass and a straightedge does not mean that it is impossible to trisect all angles: for example,
it is relatively straightforward to trisect a right angle (that is, to construct an angle of measure 30
degrees).

It is, however, possible to trisect an arbitrary angle, but using tools other than straightedge and
compass. For example, neusis construction, also known to ancient Greeks, involves simultaneous
sliding and rotation of a marked straightedge, which can not be achieved with the original tools.
Other techniques were developed by mathematicians over centuries.
Because it is defined in simple terms, but complex to prove unsolvable, the problem of angle
trisection is a frequent subject of pseudomathematical attempts at solution by naive enthusiasts.
The "solutions" often involve finding loopholes in the rules, or are simply incorrect.[1]

Contents
1 Background and problem statement

2 Proof of impossibility

3 Angles which can be trisected

o 3.1 One general theorem

4 Trisection using other methods

o 4.1 By infinite repetition of bisection

o 4.2 Using origami

o 4.3 With an auxiliary curve

o 4.4 With a marked ruler

o 4.5 With a string

o 4.6 With a "tomahawk"

o 4.7 With interconnected compasses

5 See also

6 References

7 Additional references

8 External links

o 8.1 Other means of trisection

Background and problem statement


Bisection of arbitrary angles has long been solved.

Using only an unmarked straightedge and a compass, Greek mathematicians found means to
divide a line into an arbitrary set of equal segments, to draw parallel lines, to bisect angles, to
construct many polygons, and to construct squares of equal or twice the area of a given polygon.

Three problems proved elusive, specifically, trisecting the angle, doubling the cube, and squaring
the circle. The problem of angle trisection reads:

Construct an angle equal to one-third of a given arbitrary angle (or divide it into three equal
angles), using only two tools:

1. an un-marked straightedge and

2. a compass.

Proof of impossibility

Rulers. The displayed ones are marked an ideal straightedge is un-marked


compasses

The problem of constructing an angle of a given measure is equivalent to construct two


segments such the ratio of their length is because one may pass from one solution to the
other by a compass and straightedge construction. It follows that, given a segment that is sought
as having a unit length, the problem of angle trisection is equivalent to construct a segment
whose length is the root of a cubic polynomial since by the triple-angle formula,
This allows to reduce the original geometric problem
to a purely algebraic problem.

One can show that every rational number is constructible and that every irrational number which
is constructible in one step from some given numbers is a solution of a polynomial of degree 2
with coefficients in the field generated by these numbers. Therefore any number which is
constructible by a series of steps is a root of a minimal polynomial whose degree is a power of 2.
Note also that radians (60 degrees, written 60) is constructible. We now show that it is
impossible to construct a 20 angle; this implies that a 60 angle cannot be trisected, and thus
that an arbitrary angle cannot be trisected.

Denote the set of rational numbers by Q. If 60 could be trisected, the degree of a minimal
polynomial of cos(20) over Q would be a power of two. Now let y = cos(20).

Note that cos(60) . Then by the triple-angle formula,


and so . Thus ,
or equivalently . Now substitute , so that
. Let .

The minimal polynomial for x (hence cos(20)) is a factor of . Because is degree 3, if


it is reducible over by Q then it has a rational root. By the rational root theorem, this root must be
1 or 1, but both are clearly not roots. Therefore is irreducible over by Q, and the minimal
polynomial for cos(20) is of degree 3.

So an angle of 60 = (1/3) radians cannot be trisected.


Many people (who presumably are unaware of the above result, misunderstand it, or incorrectly
reject it) have proposed methods of trisecting the general angle. Some of these methods provide
reasonable approximations; others (some of which are mentioned below) involve tools not
permitted in the classical problem. The mathematician Underwood Dudley has detailed some of
these failed attempts in his book The Trisectors.[2]

Angles which can be trisected


However, some angles can be trisected. For example, for any angle , the angle can be
trivially trisected by ignoring the given angle and directly constructing an angle of measure .
There are angles which are not constructible, but are trisectible. For example, is such an
angle: five copies of combine to make an angle of measure , which is a full circle
plus the desired . More generally, for a positive integer , an angle of measure is
trisectible if and only if does not divide ;[3] if is a prime number, this angle is
constructible if and only is a Fermat prime.

One general theorem

Again, denote the rational numbers Q:

Theorem: The angle may be trisected if and only if is reducible


over the field extension Q .

The proof is a relatively straightforward generalization of the proof given above that a 60-degree
angle is not trisectible.[4]

Trisection using other methods


The general problem of angle trisection is solvable, but using additional tools, and thus going
outside of the original Greek framework of compass and straightedge.

By infinite repetition of bisection

Trisection can be achieved by infinite repetition of the compass and straightedge method for
bisecting an angle. The geometric series 1/3 = 1/4+1/16+1/64+1/256+... or 1/3 = 1/2-1/4+1/8-
1/16+... can be used as a basis for the bisections. This method is considered to be breaking the
rules for compass and straightedge construction as it involves an infinite number of steps.
However, an approximation to any degree of accuracy can be obtained in a finite number of
steps.[5]

Using origami

Main article: Mathematics of origami#Trisecting an angle


Trisection, like many constructions impossible by ruler and compass, can easily be accomplished
by the more powerful (but physically easy) operations of paper folding, or origami. Huzita's
axioms (types of folding operations) can construct cubic extensions (cube roots) of given lengths,
whereas ruler-and-compass can construct only quadratic extensions (square roots).

With an auxiliary curve

There are certain curves called trisectrices which, if drawn on the plane using other methods, can
be used to trisect arbitrary angles.[6]

With a marked ruler

Another means to trisect an arbitrary angle by a "small" step outside the Greek framework is via
a ruler with two marks a set distance apart. The next construction is originally due to
Archimedes, called a Neusis construction, i.e., that uses tools other than an un-marked
straightedge.

This requires three facts from geometry (at right):

1. Any full set of angles on a straight line add to 180,

2. The sum of angles of any triangle is 180, and,

3. Any two equal sides of an isosceles triangle will meet the third in the same angle.
Trisection of the angle using marked ruler

At the diagram at right, angle a (left of point B) is the subject of trisection. First, a point A is
drawn at an angle's ray, one unit apart from B. A circle of radius AB is drawn.

Then, the markedness of the ruler comes into play: it is "anchored" at point A, and slided and
rotated until one mark is at point C, and one at point D, i.e., CD = AB. A radius BC is drawn as
obvious. That is to say, line segments AB, BC, and CD all have equal length. (Segment AC is
irrelevant.) Now, Triangles ABC and BCD are isosceles, thus (by Fact 3 above) each has two
equal angles.

Hypothesis: Given AD is a straight line, and AB, BC, and CD are all equal length,

Conclusion: angle .

Proof:

1. From Fact 1) above, .

2. Looking at triangle BCD, from Fact 2) .

3. From the last two equations, .

4. From Fact 2), , thus , so from last, .

5. From Fact 1) above, , thus .

Clearing, , or , and the theorem is proved.

Again, this construction stepped outside the framework of allowed constructions by using a
marked straightedge.

With a string

Thomas Hutcheson published an article in the Mathematics Teacher[7] that used a string instead
of a compass and straight edge. A string can be used as either a straight edge (by stretching it) or
a compass (by fixing one point and identifying another), but can also wrap around a cylinder, the
key to Hutcheson's solution.

Hutcheson constructed a cylinder from the angle to be trisected by drawing an arc across the
angle, completing it as a circle, and constructing from that circle a cylinder on which a, say,
equilateral triangle was inscribed (a 360-degree angle divided in three). This was then "mapped"
onto the angle to be trisected, with a simple proof of similar triangles.
With a "tomahawk"

A tomahawk trisecting an angle. The handle forms one trisector and the blue line shown forms
the other.

A "tomahawk" is a geometric shape consisting of a semicircle and two orthogonal line segments,
such that the length of the shorter segment is equal to the circle radius. Trisection is executed by
leaning the end of the tomahawk's shorter segment on one ray, the circle's edge on the other, so
that the "handle" (longer segment) crosses the angle's vertex; the trisection line runs between the
vertex and the center of the semicircle.

Note that while a tomahawk is constructible with compass and straightedge, it is not generally
possible to construct a tomahawk in any desired position. Thus, the above construction does not
contradict the nontrisectibility of angles with ruler and compass alone.

With interconnected compasses

An angle can be trisected with a device that is essentially a four-pronged version of a compass,
with linkages between the prongs designed to keep the three angles between adjacent prongs
equal.[8]

Doubling the cube


From Wikipedia, the free encyclopedia
Jump to: navigation, search

Doubling the cube (also known as the Delian problem) is one of the three most famous
geometric problems unsolvable by compass and straightedge construction. It was known to the
Egyptians, Greeks, and Indians.[1]
To "double the cube" means to be given a cube of some side length s and volume V= s3, and to
construct the side of a new cube, larger than the first, with volume 2V and therefore side length
. The problem is known to be impossible to solve with only compass and straightedge,
because 1.25992105 is not a constructible number.

Contents
1 History

2 Solutions

o 2.1 Using a marked ruler

3 References

4 External links

History
The problem owes its name to a story concerning the citizens of Delos, who consulted the oracle
at Delphi in order to learn how to defeat a plague sent by Apollo.[2] According to Plutarch[3] it
was the citizens of Delos who consulted the oracle at Delphi, seeking a solution for their internal
political problems at the time, which had intensified relationships among the citizens. The oracle
responded that they must double the size of the altar to Apollo, which was a regular cube. The
answer seemed strange to the Delians and they consulted Plato, who was able to interpret the
oracle as the mathematical problem of doubling the volume of a given cube, thus explaining the
oracle as the advice of Apollo for the citizens of Delos to occupy themselves with the study of
geometry and mathematics in order to calm down their passions.[4]

According to Plutarch, Plato gave the problem to Eudoxus and Archytas and Menaechmus, who
solved the problem using mechanical means, earning a rebuke from Plato for not solving the
problem using pure geometry (Plut., Quaestiones convivales VIII.ii, 718ef). This may be why the
problem is referred to in the 350s BC by the author of the pseudo-Platonic Sisyphus (388e) as
still unsolved.[5] However another version of the story says that all three found solutions but they
were too abstract to be of practical value[citation needed].

A significant development in finding a solution to the problem was the discovery by Hippocrates
of Chios that it is equivalent to finding two mean proportionals between a line segment and
another with twice the length.[6] In modern notation, this means that given segments of lengths a
and 2a, the duplication of the cube is equivalent to finding segments of lengths r and s so that
In turn, this means that

But Pierre Wantzel proved in 1837 that the cube root of 2 is not constructible; that is, it cannot be
constructed with straightedge and compass.

Solutions
Menaechmus' original solution involves the intersection of two conic curves. Other more
complicated methods of doubling the cube involve the cissoid of Diocles, the conchoid of
Nicomedes, or the Philo line. Archytas solved the problem in the fourth century B.C. using
geometric construction in three dimensions, determining a certain point as the intersection of
three surfaces of revolution.

False claims of doubling the cube with compass and straightedge abound in mathematical crank
literature (pseudomathematics).

Origami may also be used to construct the cube root of two by folding paper.

Using a marked ruler

There is a simple neusis construction using a marked ruler for a length which is the cube root of
2 times another length.[7]

Mark a ruler with the given length, this will eventually be GH.

Construct an equilateral triangle with the given length as side.

Extend AB an equal amount again to D.

Extend the line BC forming the line CE.

Extend the line DC forming the line CF


Place the marked ruler so it goes through A and one end G of the marked length falls on
CF and the other end of the marked length falls on ray CE. Thus GH is the given length.

The AG is the given length times the cube root of 2.

Bisection
From Wikipedia, the free encyclopedia
Jump to: navigation, search
For the bisection theorem in measure theory, see ham sandwich theorem. For other uses, see
bisect.

Line DE bisects line AB at D, line EF is a perpendicular bisector of segment AD at C, and line


EF is the interior bisector of right angle AED

In geometry, bisection is the division of something into two equal or congruent parts, usually by
a line, which is then called a bisector. The most often considered types of bisectors are the
segment bisector (a line that passes through the midpoint of a given segment) and the angle
bisector (a line that passes through the apex of an angle, that divides it into two equal angles).

In three dimensional space, bisection is usually done by a plane, also called the bisector or
bisecting plane.

Contents
1 Line segment bisector

2 Angle bisector

o 2.1 Angle bisectors of a triangle

o 2.2 Angle bisectors of a rhombus

3 Area bisectors and area-perimeter bisectors of a triangle


4 Area and diagonal bisectors of a parallelogram

5 Angle bisector theorem

6 References

7 External links

Line segment bisector

A line segment bisector passes through the midpoint of the segment. Particularly important is the
perpendicular bisector of a segment, which, according to its name, meets the segment at right
angles. The perpendicular bisector of a segment also has the property that each of its points is
equidistant from the segment's endpoints. Therefore Voronoi diagram boundaries consist of
segments of such lines or planes.

In classical geometry, the bisection is a simple compass and straightedge, whose possibility
depends on the ability to draw circles of equal radii and different centers. The segment is
bisected by drawing intersecting circles of equal radius, whose centers are the endpoints of the
segment and such that each circle goes through one endpoint. The line determined by the points
of intersection of the two circles is the perpendicular bisector of the segment, since it crosses the
segment at its center. This construction is in fact used when constructing a line perpendicular to a
given line at a given point: drawing an arbitrary circle whose center is that point, it intersects the
line in two more points, and the perpendicular to be constructed is the one bisecting the segment
defined by these two points.

Angle bisector
Bisection of an angle using a compass and ruler

An angle bisector divides the angle into two angles with equal measures. An angle only has one
bisector. Each point of an angle bisector is equidistant from the sides of the angle.

The interior bisector of an angle is the half-line or line segment that divides an angle of less than
180 into two equal angles. The exterior bisector is the half-line that divides the opposite angle
(of greater than 180) into two equal angles.

To bisect an angle with straightedge and compass, one draws a circle whose center is the vertex.
The circle meets the angle at two points: one on each leg. Using each of these points as a center,
draw two circles of the same size. The intersection of the circles (two points) determines a line
that is the angle bisector.

The proof of the correctness of these two constructions is fairly intuitive, relying on the
symmetry of the problem. It is interesting to note that the trisection of an angle (dividing it into
three equal parts) cannot be achieved with the compass and ruler alone (this was first proved by
Pierre Wantzel).

Angle bisectors of a triangle

The angle bisectors of the angles of a triangle are concurrent in a point called the incenter of the
triangle.

If the side lengths of a triangle are , the semiperimeter ( ) is , and A is


the angle opposite side , the length of the internal bisector of angle A is[1]

If the internal bisector of angle A in triangle ABC has length and if this bisector divides the
side opposite A into segments of lengths m and n, then[1]
where b and c are the side lengths opposite vertices B and C; and the side opposite A is divided
in the proportion b:c.

If the internal bisectors of angles A, B, and C have lengths and , then[2]

The angle bisector theorem is concerned with the relative lengths of the two segments that a
triangle's side is divided into by a line that bisects the opposite angle. It equates their relative
lengths to the relative lengths of the other two sides of the triangle.

No two non-congruent triangles share the same set of three internal angle bisector lengths.[3][4]

Angle bisectors of a rhombus

Each diagonal of a rhombus bisects opposite angles.

Area bisectors and area-perimeter bisectors of a triangle


There are an infinitude of lines that bisect the area of a triangle. Three of them are the medians of
the triangle (which connect the sides' midpoints with the opposite vertices), and these are
concurrent at the triangle's centroid; indeed, they are the only area bisectors that go through the
centroid. Three other area bisectors are parallel to the triangle's sides; each of these intersects the
other two sides so as to divide them into segments with the proportions .[5] These six
lines are concurrent three at a time: in addition to the three medians being concurrent, any one
median is concurrent with two of the side-parallel area bisectors.

The envelope of the infinitude of area bisectors is a deltoid (broadly defined as a figure with
three vertices connected by curves that are concave to the exterior of the deltoid, making the
interior points a non-convex set).[5] The vertices of the deltoid are at the midpoints of the
medians; all points inside the deltoid are on three different area bisectors, while all points outside
it are on just one. [1] The sides of the deltoid are arcs of hyperbolas that are asymptotic to the
extended sides of the triangle.[5]

Any line through a triangle that splits both the triangle's area and its perimeter in half goes
through the triangle's incenter (the center of its incircle). There are either one, two, or three of
these for any given triangle. A line through the incenter bisects one of the area or perimeter if and
only if it also bisects the other.[6]

Area and diagonal bisectors of a parallelogram


Any line through the midpoint of a parallelogram bisects the area.[7] Also, the diagonals of a
parallelogram bisect each other.

Angle bisector theorem


Main article: Angle bisector theorem

In this diagram, BD:DC = AB:AC.

In geometry, the angle bisector theorem is concerned with the relative lengths of the two
segments that a triangle's side is divided into by a line that bisects the opposite angle. It equates
their relative lengths to the relative lengths of the other two sides of the triangle.

Euclidean geometry
From Wikipedia, the free encyclopedia
Jump to: navigation, search
"Plane geometry" redirects here. For other uses, see Plane geometry (disambiguation).
A Greek mathematician performing a geometric construction with a compass, from The School
of Athens by Raphael
Geometry

Oxyrhynchus papyrus (P.Oxy. I 29) showing fragment of


Euclid's Elements

History of geometry

Branches[show]

Research areas[show]

Important concepts[show]

Geometers[show]

v
t

Euclidean geometry is a mathematical system attributed to the Alexandrian Greek


mathematician Euclid, which he described in his textbook on geometry: the Elements. Euclid's
method consists in assuming a small set of intuitively appealing axioms, and deducing many
other propositions (theorems) from these. Although many of Euclid's results had been stated by
earlier mathematicians,[1] Euclid was the first to show how these propositions could fit into a
comprehensive deductive and logical system.[2] The Elements begins with plane geometry, still
taught in secondary school as the first axiomatic system and the first examples of formal proof. It
goes on to the solid geometry of three dimensions. Much of the Elements states results of what
are now called algebra and number theory, couched in geometrical language.[3]

For over two thousand years, the adjective "Euclidean" was unnecessary because no other sort of
geometry had been conceived. Euclid's axioms seemed so intuitively obvious (with the possible
exception of the parallel postulate) that any theorem proved from them was deemed true in an
absolute, often metaphysical, sense. Today, however, many other self-consistent non-Euclidean
geometries are known, the first ones having been discovered in the early 19th century. An
implication of Einstein's theory of general relativity is that Euclidean space is a good
approximation to the properties of physical space only where the gravitational field is weak.[4]

Contents
1 The Elements

o 1.1 Axioms

o 1.2 Parallel postulate

2 Methods of proof

3 System of measurement and arithmetic

4 Notation and terminology

o 4.1 Naming of points and figures


o 4.2 Complementary and supplementary angles

o 4.3 Modern versions of Euclid's notation

5 Some important or well known results

o 5.1 Bridge of Asses

o 5.2 Congruence of triangles

o 5.3 Sum of the angles of a triangle acute, obtuse, and right angle limits

o 5.4 Pythagorean theorem

o 5.5 Thales' theorem

o 5.6 Scaling of area and volume

6 Applications

7 As a description of the structure of space

8 Later work

o 8.1 Archimedes and Apollonius

o 8.2 17th century: Descartes

o 8.3 18th century

o 8.4 19th century and non-Euclidean geometry

o 8.5 20th century and general relativity

9 Treatment of infinity

o 9.1 Infinite objects

o 9.2 Infinite processes

10 Logical basis
o 10.1 Classical logic

o 10.2 Modern standards of rigor

o 10.3 Axiomatic formulations

o 10.4 Constructive approaches and pedagogy

11 See also

o 11.1 Classical theorems

12 Notes

13 References

14 External links

The Elements
Main article: Euclid's Elements

The Elements are mainly a systematization of earlier knowledge of geometry. Its superiority over
earlier treatments was rapidly recognized, with the result that there was little interest in
preserving the earlier ones, and they are now nearly all lost.

Books IIV and VI discuss plane geometry. Many results about plane figures are proved, e.g., If
a triangle has two equal angles, then the sides subtended by the angles are equal. The
Pythagorean theorem is proved.[5]

Books V and VIIX deal with number theory, with numbers treated geometrically via their
representation as line segments with various lengths. Notions such as prime numbers and rational
and irrational numbers are introduced. The infinitude of prime numbers is proved.

Books XIXIII concern solid geometry. A typical result is the 1:3 ratio between the volume of a
cone and a cylinder with the same height and base.
The parallel postulate: If two lines intersect a third in such a way that the sum of the inner angles
on one side is less than two right angles, then the two lines inevitably must intersect each other
on that side if extended far enough.

Axioms

Euclidean geometry is an axiomatic system, in which all theorems ("true statements") are derived
from a small number of axioms.[6] Near the beginning of the first book of the Elements, Euclid
gives five postulates (axioms) for plane geometry, stated in terms of constructions (as translated
by Thomas Heath):[7]

"Let the following be postulated":

1. "To draw a straight line from any point to any point."

2. "To produce [extend] a finite straight line continuously in a straight line."

3. "To describe a circle with any centre and distance [radius]."

4. "That all right angles are equal to one another."

5. The parallel postulate: "That, if a straight line falling on two straight lines make the
interior angles on the same side less than two right angles, the two straight lines, if
produced indefinitely, meet on that side on which are the angles less than the two right
angles."

Although Euclid's statement of the postulates only explicitly asserts the existence of the
constructions, they are also taken to be unique.

The Elements also include the following five "common notions":

1. Things that are equal to the same thing are also equal to one another (Transitive property
of equality).

2. If equals are added to equals, then the wholes are equal.


3. If equals are subtracted from equals, then the remainders are equal.

4. Things that coincide with one another equal one another (Reflexive Property).

5. The whole is greater than the part.

Parallel postulate

Main article: Parallel postulate

To the ancients, the parallel postulate seemed less obvious than the others. They were concerned
with creating a system which was absolutely rigorous and to them it seemed as if the parallel line
postulate should have been able to be proven rather than simply accepted as a fact. It is now
known that such a proof is impossible. Euclid himself seems to have considered it as being
qualitatively different from the others, as evidenced by the organization of the Elements: the first
28 propositions he presents are those that can be proved without it.

Many alternative axioms can be formulated that have the same logical consequences as the
parallel postulate. For example Playfair's axiom states:

In a plane, through a point not on a given straight line, at most one line can be drawn that
never meets the given line.

A proof from Euclid's elements that, given a line segment, an equilateral triangle exists that
includes the segment as one of its sides. The proof is by construction: an equilateral triangle
is made by drawing circles and centered on the points and , and taking one intersection of
the circles as the third vertex of the triangle.

Methods of proof
Euclidean Geometry is constructive. Postulates 1, 2, 3, and 5 assert the existence and uniqueness
of certain geometric figures, and these assertions are of a constructive nature: that is, we are not
only told that certain things exist, but are also given methods for creating them with no more
than a compass and an unmarked straightedge.[8] In this sense, Euclidean geometry is more
concrete than many modern axiomatic systems such as set theory, which often assert the
existence of objects without saying how to construct them, or even assert the existence of objects
that cannot be constructed within the theory.[9] Strictly speaking, the lines on paper are models of
the objects defined within the formal system, rather than instances of those objects. For example
a Euclidean straight line has no width, but any real drawn line will. Though nearly all modern
mathematicians consider nonconstructive methods just as sound as constructive ones, Euclid's
constructive proofs often supplanted fallacious nonconstructive onese.g., some of the
Pythagoreans' proofs that involved irrational numbers, which usually required a statement such
as "Find the greatest common measure of ..."[10]

Euclid often used proof by contradiction. Euclidean geometry also allows the method of
superposition, in which a figure is transferred to another point in space. For example, proposition
I.4, side-angle-side congruence of triangles, is proved by moving one of the two triangles so that
one of its sides coincides with the other triangle's equal side, and then proving that the other
sides coincide as well. Some modern treatments add a sixth postulate, the rigidity of the triangle,
which can be used as an alternative to superposition.[11]

System of measurement and arithmetic


Euclidean geometry has two fundamental types of measurements: angle and distance. The angle
scale is absolute, and Euclid uses the right angle as his basic unit, so that, e.g., a 45-degree angle
would be referred to as half of a right angle. The distance scale is relative; one arbitrarily picks a
line segment with a certain length as the unit, and other distances are expressed in relation to it.

A line in Euclidean geometry is a model of the real number line. A line segment is a part of a line
that is bounded by two end points, and contains every point on the line between its end points.
Addition is represented by a construction in which one line segment is copied onto the end of
another line segment to extend its length, and similarly for subtraction.

Measurements of area and volume are derived from distances. For example, a rectangle with a
width of 3 and a length of 4 has an area that represents the product, 12. Because this geometrical
interpretation of multiplication was limited to three dimensions (as to multiply three numbers in
a Euclidean interpretation, a 3 dimensional rectangular prism would have to have been created,
and if one would have wanted to a greater amount of numbers, they would have to move into
higher dimensions, which Euclid did not accept the existence of[citation needed]), there was no direct
way of interpreting the product of four or more numbers, and Euclid avoided such products,
although they are implied, e.g., in the proof of book IX, proposition 20.
An example of congruence. The two figures on the left are congruent, while the third is similar to
them. The last figure is neither. Note that congruences alter some properties, such as location and
orientation, but leave others unchanged, like distance and angles. The latter sort of properties are
called invariants and studying them is the essence of geometry.

Euclid refers to a pair of lines, or a pair of planar or solid figures, as "equal" () if their
lengths, areas, or volumes are equal, and similarly for angles. The stronger term "congruent"
refers to the idea that an entire figure is the same size and shape as another figure. Alternatively,
two figures are congruent if one can be moved on top of the other so that it matches up with it
exactly. (Flipping it over is allowed.) Thus, for example, a 2x6 rectangle and a 3x4 rectangle are
equal but not congruent, and the letter R is congruent to its mirror image. Figures that would be
congruent except for their differing sizes are referred to as similar. Corresponding angles in a
pair of similar shapes are congruent and corresponding sides are in proportion to each other.

Notation and terminology


Naming of points and figures

Points are customarily named using capital letters of the alphabet. Other figures, such as lines,
triangles, or circles, are named by listing a sufficient number of points to pick them out
unambiguously from the relevant figure, e.g., triangle ABC would typically be a triangle with
vertices at points A, B, and C.

Complementary and supplementary angles

Angles whose sum is a right angle are called complementary. Complementary angles are formed
when one or more rays share the same vertex and are pointed in a direction that is in between the
two original rays that form the right angle. The number of rays in between the two original rays
are infinite. Those whose sum is a straight angle are supplementary. Supplementary angles are
formed when one or more rays share the same vertex and are pointed in a direction that in
between the two original rays that form the straight angle (180 degrees). The number of rays in
between the two original rays are infinite like those possible in the complementary angle.

Modern versions of Euclid's notation

In modern terminology, angles would normally be measured in degrees or radians.

Modern school textbooks often define separate figures called lines (infinite), rays (semi-infinite),
and line segments (of finite length). Euclid, rather than discussing a ray as an object that extends
to infinity in one direction, would normally use locutions such as "if the line is extended to a
sufficient length," although he occasionally referred to "infinite lines." A "line" in Euclid could
be either straight or curved, and he used the more specific term "straight line" when necessary.

Some important or well known results


The bridge of asses theorem states that if A=B then C=D.

The sum of angles A, B, and C is equal to 180 degrees.

Pythagoras' theorem: The sum of the areas of the two squares on the legs (a and b) of a
right triangle equals the area of the square on the hypotenuse (c).

Thales' theorem: if AC is a diameter, then the angle at B is a right angle.

Bridge of Asses

The Bridge of Asses (Pons Asinorum) states that in isosceles triangles the angles at the base
equal one another, and, if the equal straight lines are produced further, then the angles under the
base equal one another.[12] Its name may be attributed to its frequent role as the first real test in
the Elements of the intelligence of the reader and as a bridge to the harder propositions that
followed. It might also be so named because of the geometrical figure's resemblance to a steep
bridge that only a sure-footed donkey could cross.[13]

Congruence of triangles
Congruence of triangles is determined by specifying two sides and the angle between them
(SAS), two angles and the side between them (ASA) or two angles and a corresponding adjacent
side (AAS). Specifying two sides and an adjacent angle (SSA), however, can yield two distinct
possible triangles.

Triangles are congruent if they have all three sides equal (SSS), two sides and the angle between
them equal (SAS), or two angles and a side equal (ASA) (Book I, propositions 4, 8, and 26).
(Triangles with three equal angles (AAA) are similar, but not necessarily congruent. Also,
triangles with two equal sides and an adjacent angle are not necessarily equal or congruent.)

Sum of the angles of a triangle acute, obtuse, and right angle limits

The sum of the angles of a triangle is equal to a straight angle (180 degrees).[14] This causes an
equilateral triangle to have 3 interior angles of 60 degrees. Also, it causes every triangle to have
at least 2 acute angles and up to 1 obtuse or right angle.

Pythagorean theorem

The celebrated Pythagorean theorem (book I, proposition 47) states that in any right triangle, the
area of the square whose side is the hypotenuse (the side opposite the right angle) is equal to the
sum of the areas of the squares whose sides are the two legs (the two sides that meet at a right
angle).

Thales' theorem

Thales' theorem, named after Thales of Miletus states that if A, B, and C are points on a circle
where the line AC is a diameter of the circle, then the angle ABC is a right angle. Cantor
supposed that Thales proved his theorem by means of Euclid book I, prop 32 after the manner of
Euclid book III, prop 31.[15] Tradition has it that Thales sacrificed an ox to celebrate this theorem.
[16]
Scaling of area and volume

In modern terminology, the area of a plane figure is proportional to the square of any of its linear
dimensions, , and the volume of a solid to the cube, . Euclid proved these
results in various special cases such as the area of a circle[17] and the volume of a parallelepipedal
solid.[18] Euclid determined some, but not all, of the relevant constants of proportionality. E.g., it
was his successor Archimedes who proved that a sphere has 2/3 the volume of the
circumscribing cylinder.[19]

Applications
This section requires expansion. (March 2009)

Because of Euclidean geometry's fundamental status in mathematics, it would be impossible to


give more than a representative sampling of applications here.

A surveyor uses a Level

Sphere packing applies to a stack of oranges.

A parabolic mirror brings parallel rays of light to a focus.

As suggested by the etymology of the word, one of the earliest reasons for interest in geometry
was surveying,[20] and certain practical results from Euclidean geometry, such as the right-angle
property of the 3-4-5 triangle, were used long before they were proved formally.[21] The
fundamental types of measurements in Euclidean geometry are distances and angles, and both of
these quantities can be measured directly by a surveyor. Historically, distances were often
measured by chains such as Gunter's chain, and angles using graduated circles and, later, the
theodolite.

An application of Euclidean solid geometry is the determination of packing arrangements, such


as the problem of finding the most efficient packing of spheres in n dimensions. This problem
has applications in error detection and correction.

Geometric optics uses Euclidean geometry to analyze the focusing of light by lenses and mirrors.

Geometry is used in art and architecture.

The water tower consists of a cone, a cylinder, and a hemisphere. Its volume can be
calculated using solid geometry.

Geometry can be used to design origami.

Geometry is used extensively in architecture.

Geometry can be used to design origami. Some classical construction problems of geometry are
impossible using compass and straightedge, but can be solved using origami.[22]

As a description of the structure of space


Euclid believed that his axioms were self-evident statements about physical reality. Euclid's
proofs depend upon assumptions perhaps not obvious in Euclid's fundamental axioms,[23] in
particular that certain movements of figures do not change their geometrical properties such as
the lengths of sides and interior angles, the so-called Euclidean motions, which include
translations and rotations of figures.[24] Taken as a physical description of space, postulate 2
(extending a line) asserts that space does not have holes or boundaries (in other words, space is
homogeneous and unbounded); postulate 4 (equality of right angles) says that space is isotropic
and figures may be moved to any location while maintaining congruence; and postulate 5 (the
parallel postulate) that space is flat (has no intrinsic curvature).[25]

As discussed in more detail below, Einstein's theory of relativity significantly modifies this view.

The ambiguous character of the axioms as originally formulated by Euclid makes it possible for
different commentators to disagree about some of their other implications for the structure of
space, such as whether or not it is infinite[26] (see below) and what its topology is. Modern, more
rigorous reformulations of the system[27] typically aim for a cleaner separation of these issues.
Interpreting Euclid's axioms in the spirit of this more modern approach, axioms 1-4 are
consistent with either infinite or finite space (as in elliptic geometry), and all five axioms are
consistent with a variety of topologies (e.g., a plane, a cylinder, or a torus for two-dimensional
Euclidean geometry).

Later work
Archimedes and Apollonius

A sphere has 2/3 the volume and surface area of its circumscribing cylinder. A sphere and
cylinder were placed on the tomb of Archimedes at his request.

Archimedes (ca. 287 BCE ca. 212 BCE), a colorful figure about whom many historical
anecdotes are recorded, is remembered along with Euclid as one of the greatest of ancient
mathematicians. Although the foundations of his work were put in place by Euclid, his work,
unlike Euclid's, is believed to have been entirely original.[28] He proved equations for the volumes
and areas of various figures in two and three dimensions, and enunciated the Archimedean
property of finite numbers.

Apollonius of Perga (ca. 262 BCEca. 190 BCE) is mainly known for his investigation of conic
sections.

Ren Descartes. Portrait after Frans Hals, 1648.

17th century: Descartes

Ren Descartes (15961650) developed analytic geometry, an alternative method for formalizing
geometry which focused on turning geometry into algebra.[29] In this approach, a point is
represented by its Cartesian (x, y) coordinates, a line is represented by its equation, and so on. In
Euclid's original approach, the Pythagorean theorem follows from Euclid's axioms. In the
Cartesian approach, the axioms are the axioms of algebra, and the equation expressing the
Pythagorean theorem is then a definition of one of the terms in Euclid's axioms, which are now
considered theorems. The equation

defining the distance between two points P = (p, q) and Q = (r, s) is then known as the Euclidean
metric, and other metrics define non-Euclidean geometries.

In terms of analytic geometry, the restriction of classical geometry to compass and straightedge
constructions means a restriction to first- and second-order equations, e.g., y = 2x + 1 (a line), or
x2 + y2 = 7 (a circle).

Also in the 17th century, Girard Desargues, motivated by the theory of perspective, introduced
the concept of idealized points, lines, and planes at infinity. The result can be considered as a
type of generalized geometry, projective geometry, but it can also be used to produce proofs in
ordinary Euclidean geometry in which the number of special cases is reduced.[30]

Squaring the circle: the areas of this square and this circle are equal. In 1882, it was proven that
this figure cannot be constructed in a finite number of steps with an idealized compass and
straightedge.

18th century

Geometers of the 18th century struggled to define the boundaries of the Euclidean system. Many
tried in vain to prove the fifth postulate from the first four. By 1763 at least 28 different proofs
had been published, but all were found incorrect.[31]

Leading up to this period, geometers also tried to determine what constructions could be
accomplished in Euclidean geometry. For example, the problem of trisecting an angle with a
compass and straightedge is one that naturally occurs within the theory, since the axioms refer to
constructive operations that can be carried out with those tools. However, centuries of efforts
failed to find a solution to this problem, until Pierre Wantzel published a proof in 1837 that such
a construction was impossible. Other constructions that were proved impossible include doubling
the cube and squaring the circle. In the case of doubling the cube, the impossibility of the
construction originates from the fact that the compass and straightedge method involve first- and
second-order equations, while doubling a cube requires the solution of a third-order equation.

Euler discussed a generalization of Euclidean geometry called affine geometry, which retains the
fifth postulate unmodified while weakening postulates three and four in a way that eliminates the
notions of angle (whence right triangles become meaningless) and of equality of length of line
segments in general (whence circles become meaningless) while retaining the notions of
parallelism as an equivalence relation between lines, and equality of length of parallel line
segments (so line segments continue to have a midpoint).

19th century and non-Euclidean geometry


This time period was not only concerned with exploring the geometry that is created by looking
at Non-Euclidean geometries, but also attempting to completely ensure that what Euclid said was
true was actually true.

In the early 19th century, Carnot and Mbius systematically developed the use of signed angles
and line segments as a way of simplifying and unifying results.[32]

The century's most significant development in geometry occurred when, around 1830, Jnos
Bolyai and Nikolai Ivanovich Lobachevsky separately published work on non-Euclidean
geometry, in which the parallel postulate is not valid.[33] Since non-Euclidean geometry is
provably relatively consistent with Euclidean geometry, the parallel postulate cannot be proved
from the other postulates.

In the 19th century, it was also realized that Euclid's ten axioms and common notions do not
suffice to prove all of theorems stated in the Elements. For example, Euclid assumed implicitly
that any line contains at least two points, but this assumption cannot be proved from the other
axioms, and therefore must be an axiom itself. The very first geometric proof in the Elements,
shown in the figure above, is that any line segment is part of a triangle; Euclid constructs this in
the usual way, by drawing circles around both endpoints and taking their intersection as the third
vertex. His axioms, however, do not guarantee that the circles actually intersect, because they do
not assert the geometrical property of continuity, which in Cartesian terms is equivalent to the
completeness property of the real numbers. Starting with Moritz Pasch in 1882, many improved
axiomatic systems for geometry have been proposed, the best known being those of Hilbert,[34]
George Birkhoff,[35] and Tarski.[36]

20th century and general relativity

A disproof of Euclidean geometry as a description of physical space. In a 1919 test of the general
theory of relativity, stars (marked with short horizontal lines) were photographed during a solar
eclipse. The rays of starlight were bent by the Sun's gravity on their way to the earth. This is
interpreted as evidence in favor of Einstein's prediction that gravity would cause deviations from
Euclidean geometry.

Einstein's theory of general relativity shows that the true geometry of spacetime is not Euclidean
geometry.[37] For example, if a triangle is constructed out of three rays of light, then in general the
interior angles do not add up to 180 degrees due to gravity. A relatively weak gravitational field,
such as the Earth's or the sun's, is represented by a metric that is approximately, but not exactly,
Euclidean. Until the 20th century, there was no technology capable of detecting the deviations
from Euclidean geometry, but Einstein predicted that such deviations would exist. They were
later verified by observations such as the slight bending of starlight by the Sun during a solar
eclipse in 1919, and such considerations are now an integral part of the software that runs the
GPS system.[38] It is possible to object to this interpretation of general relativity on the grounds
that light rays might be improper physical models of Euclid's lines, or that relativity could be
rephrased so as to avoid the geometrical interpretations. However, one of the consequences of
Einstein's theory is that there is no possible physical test that can distinguish between a beam of
light as a model of a geometrical line and any other physical model. Thus, the only logical
possibilities are to accept non-Euclidean geometry as physically real, or to reject the entire
notion of physical tests of the axioms of geometry, which can then be imagined as a formal
system without any intrinsic real-world meaning.

Treatment of infinity
Infinite objects

Euclid sometimes distinguished explicitly between "finite lines" (e.g., Postulate 2) and "infinite
lines" (book I, proposition 12). However, he typically did not make such distinctions unless they
were necessary. The postulates do not explicitly refer to infinite lines, although for example some
commentators interpret postulate 3, existence of a circle with any radius, as implying that space
is infinite.[26]

The notion of infinitesimally small quantities had previously been discussed extensively by the
Eleatic School, but nobody had been able to put them on a firm logical basis, with paradoxes
such as Zeno's paradox occurring that had not been resolved to universal satisfaction. Euclid
used the method of exhaustion rather than infinitesimals.[39]

Later ancient commentators such as Proclus (410485 CE) treated many questions about infinity
as issues demanding proof and, e.g., Proclus claimed to prove the infinite divisibility of a line,
based on a proof by contradiction in which he considered the cases of even and odd numbers of
points constituting it.[40]

At the turn of the 20th century, Otto Stolz, Paul du Bois-Reymond, Giuseppe Veronese, and
others produced controversial work on non-Archimedean models of Euclidean geometry, in
which the distance between two points may be infinite or infinitesimal, in the NewtonLeibniz
sense.[41] Fifty years later, Abraham Robinson provided a rigorous logical foundation for
Veronese's work.[42]
Infinite processes

One reason that the ancients treated the parallel postulate as less certain than the others is that
verifying it physically would require us to inspect two lines to check that they never intersected,
even at some very distant point, and this inspection could potentially take an infinite amount of
time.[43]

The modern formulation of proof by induction was not developed until the 17th century, but
some later commentators consider it implicit in some of Euclid's proofs, e.g., the proof of the
infinitude of primes.[44]

Supposed paradoxes involving infinite series, such as Zeno's paradox, predated Euclid. Euclid
avoided such discussions, giving, for example, the expression for the partial sums of the
geometric series in IX.35 without commenting on the possibility of letting the number of terms
become infinite.

Logical basis
This article needs attention from an expert in mathematics. Please add a reason or a
talk parameter to this template to explain the issue with the article. WikiProject
Mathematics or the Mathematics Portal may be able to help recruit an expert. (December
2010)
This section requires expansion. (June 2010)
See also: Hilbert's axioms, Axiomatic system, and Real closed field

Classical logic

Euclid frequently used the method of proof by contradiction, and therefore the traditional
presentation of Euclidean geometry assumes classical logic, in which every proposition is either
true or false, i.e., for any proposition P, the proposition "P or not P" is automatically true.

Modern standards of rigor

Placing Euclidean geometry on a solid axiomatic basis was a preoccupation of mathematicians


for centuries.[45] The role of primitive notions, or undefined concepts, was clearly put forward by
Alessandro Padoa of the Peano delegation at the 1900 Paris conference:[45][46]

...when we begin to formulate the theory, we can imagine that the undefined symbols are
completely devoid of meaning and that the unproved propositions are simply conditions imposed
upon the undefined symbols.

Then, the system of ideas that we have initially chosen is simply one interpretation of the
undefined symbols; but..this interpretation can be ignored by the reader, who is free to replace it
in his mind by another interpretation.. that satisfies the conditions...
Logical questions thus become completely independent of empirical or psychological
questions...

The system of undefined symbols can then be regarded as the abstraction obtained from the
specialized theories that result when...the system of undefined symbols is successively replaced
by each of the interpretations...
Padoa, Essai d'une thorie algbrique des nombre entiers, avec une Introduction logique
une thorie dductive qulelconque

That is, mathematics is context-independent knowledge within a hierarchical framework. As said


by Bertrand Russell:[47]

If our hypothesis is about anything, and not about some one or more particular things, then our
deductions constitute mathematics. Thus, mathematics may be defined as the subject in which we
never know what we are talking about, nor whether what we are saying is true.
Bertrand Russell, Mathematics and the metaphysicians

Such foundational approaches range between foundationalism and formalism.

Axiomatic formulations

Geometry is the science of correct reasoning on incorrect figures.


George Poly, How to Solve It, p. 208

Euclid's axioms: In his dissertation to Trinity College, Cambridge, Bertrand Russell


summarized the changing role of Euclid's geometry in the minds of philosophers up to
that time.[48] It was a conflict between certain knowledge, independent of experiment, and
empiricism, requiring experimental input. This issue became clear as it was discovered
that the parallel postulate was not necessarily valid and its applicability was an empirical
matter, deciding whether the applicable geometry was Euclidean or non-Euclidean.

Hilbert's axioms: Hilbert's axioms had the goal of identifying a simple and complete set
of independent axioms from which the most important geometric theorems could be
deduced. The outstanding objectives were to make Euclidean geometry rigorous
(avoiding hidden assumptions) and to make clear the ramifications of the parallel
postulate.

Birkhoff's axioms: Birkhoff proposed four postulates for Euclidean geometry that can be
confirmed experimentally with scale and protractor.[49][50][51] The notions of angle and
distance become primitive concepts.[52]

Tarski's axioms: Alfred Tarski (19021983) and his students defined elementary
Euclidean geometry as the geometry that can be expressed in first-order logic and does
not depend on set theory for its logical basis,[53] in contrast to Hilbert's axioms, which
involve point sets.[54] Tarski proved that his axiomatic formulation of elementary
Euclidean geometry is consistent and complete in a certain sense: there is an algorithm
that, for every proposition, can be shown either true or false.[36] (This doesn't violate
Gdel's theorem, because Euclidean geometry cannot describe a sufficient amount of
arithmetic for the theorem to apply.[55]) This is equivalent to the decidability of real closed
fields, of which elementary Euclidean geometry is a model.

Constructive approaches and pedagogy

The process of abstract axiomatization as exemplified by Hilbert's axioms reduces geometry to


theorem proving or predicate logic. In contrast, the Greeks used construction postulates, and
emphasized problem solving.[56] For the Greeks, constructions are more primitive than existence
propositions, and can be used to prove existence propositions, but not vice versa. To describe
problem solving adequately requires a richer system of logical concepts.[56] The contrast in
approach may be summarized:[57]

Axiomatic proof: Proofs are deductive derivations of propositions from primitive


premises that are true in some sense. The aim is to justify the proposition.

Analytic proof: Proofs are non-deductive derivations of hypotheses from problems. The
aim is to find hypotheses capable of giving a solution to the problem. One can argue that
Euclid's axioms were arrived upon in this manner. In particular, it is thought that Euclid
felt the parallel postulate was forced upon him, as indicated by his reluctance to make use
of it,[58] and his arrival upon it by the method of contradiction.[59]

Andrei Nicholaevich Kolmogorov proposed a problem solving basis for geometry.[60][61] This
work was a precursor of a modern formulation in terms of constructive type theory.[62] This
development has implications for pedagogy as well.[63]

If proof simply follows conviction of truth rather than contributing to its construction and is only
experienced as a demonstration of something already known to be true, it is likely to remain
meaningless and purposeless in the eyes of students.
Celia Hoyles, The curricular shaping of students' approach to proof

Analytic geometry
From Wikipedia, the free encyclopedia
Jump to: navigation, search
Cartesian coordinates.
Geometry

Oxyrhynchus papyrus (P.Oxy. I 29) showing fragment of


Euclid's Elements

History of geometry

Branches[show]

Research areas[show]

Important concepts[show]

Geometers[show]
v

Analytic geometry, or analytical geometry, has two different meanings in mathematics. The
modern and advanced meaning refers to the geometry of analytic varieties. This article focuses
on the classical and elementary meaning.

In classical mathematics, analytic geometry, also known as coordinate geometry, or Cartesian


geometry, is the study of geometry using a coordinate system and the principles of algebra and
analysis. This contrasts with the synthetic approach of Euclidean geometry, which treats certain
geometric notions as primitive, and uses deductive reasoning based on axioms and theorems to
derive truth. Analytic geometry is widely used in physics and engineering, and is the foundation
of most modern fields of geometry, including algebraic, differential, discrete, and computational
geometry.

Usually the Cartesian coordinate system is applied to manipulate equations for planes, straight
lines, and squares, often in two and sometimes in three dimensions. Geometrically, one studies
the Euclidean plane (2 dimensions) and Euclidean space (3 dimensions). As taught in school
books, analytic geometry can be explained more simply: it is concerned with defining and
representing geometrical shapes in a numerical way and extracting numerical information from
shapes' numerical definitions and representations. The numerical output, however, might also be
a vector or a shape. That the algebra of the real numbers can be employed to yield results about
the linear continuum of geometry relies on the CantorDedekind axiom.

Contents
1 History

2 Basic principles

o 2.1 Coordinates

o 2.2 Equations of curves

o 2.3 Distance and angle

o 2.4 Section of a line


o 2.5 Transformations

o 2.6 Intersections

o 2.7 Intercepts

3 Themes

4 Example

5 Modern analytic geometry

6 Notes

7 References

8 External links

History
The Greek mathematician Menaechmus solved problems and proved theorems by using a
method that had a strong resemblance to the use of coordinates and it has sometimes been
maintained that he had introduced analytic geometry.[1] Apollonius of Perga, in On Determinate
Section, dealt with problems in a manner that may be called an analytic geometry of one
dimension; with the question of finding points on a line that were in a ratio to the others.[2]
Apollonius in the Conics further developed a method that is so similar to analytic geometry that
his work is sometimes thought to have anticipated the work of Descartes by some 1800 years.
His application of reference lines, a diameter and a tangent is essentially no different than our
modern use of a coordinate frame, where the distances measured along the diameter from the
point of tangency are the abscissas, and the segments parallel to the tangent and intercepted
between the axis and the curve are the ordinates. He further developed relations between the
abscissas and the corresponding ordinates that are equivalent to rhetorical equations of curves.
However, although Apollonius came close to developing analytic geometry, he did not manage to
do so since he did not take into account negative magnitudes and in every case the coordinate
system was superimposed upon a given curve a posteriori instead of a priori. That is, equations
were determined by curves, but curves were not determined by equations. Coordinates, variables,
and equations were subsidiary notions applied to a specific geometric situation.[3]

The eleventh century Persian mathematician Omar Khayym saw a strong relationship between
geometry and algebra, and was moving in the right direction when he helped to close the gap
between numerical and geometric algebra[4] with his geometric solution of the general cubic
equations,[5] but the decisive step came later with Descartes.[4]
Analytic geometry has traditionally been attributed to Ren Descartes[4][6][7] Descartes made
significant progress with the methods in an essay entitled La Geometrie (Geometry), one of the
three accompanying essays (appendices) published in 1637 together with his Discourse on the
Method for Rightly Directing One's Reason and Searching for Truth in the Sciences, commonly
referred to as Discourse on Method. This work, written in his native French tongue, and its
philosophical principles, provided a foundation for Infinitesimal calculus in Europe. Initially the
work was not well received, due, in part, to the many gaps in arguments and complicated
equations. Only after the translation into Latin and the addition of commentary by van Schooten
in 1649 (and further work thereafter) did Descarte's masterpiece receive due recognition.[8]

Pierre Fermat also pioneered the development of analytic geometry. Although not published in
his lifetime, a manuscript form of Ad locos planos et solidos isagoge (Introduction to Plane and
Solid Loci) was circulating in Paris in 1637, just prior to the publication of Descartes' Discourse.
[9]
Clearly written and well received, the Introduction also laid the groundwork for analytical
geometry. The key difference between Fermat's and Descartes' treatments is a matter of
viewpoint. Fermat always started with an algebraic equation and then described the geometric
curve which satisfied it, while Descartes starts with geometric curves and produces their
equations as one of several properties of the curves.[8] As a consequence of this approach,
Descartes had to deal with more complicated equations and he had to develop the methods to
work with polynomial equations of higher degree.

Basic principles

Illustration of a Cartesian coordinate plane. Four points are marked and labeled with their
coordinates: (2,3) in green, (3,1) in red, (1.5,2.5) in blue, and the origin (0,0) in purple.

Coordinates

Main article: Coordinate systems


In analytic geometry, the plane is given a coordinate system, by which every point has a pair of
real number coordinates. The most common coordinate system to use is the Cartesian coordinate
system, where each point has an x-coordinate representing its horizontal position, and a y-
coordinate representing its vertical position. These are typically written as an ordered pair (x, y).
This system can also be used for three-dimensional geometry, where every point in Euclidean
space is represented by an ordered triple of coordinates (x, y, z).

Other coordinate systems are possible. On the plane the most common alternative is polar
coordinates, where every point is represented by its radius r from the origin and its angle . In
three dimensions, common alternative coordinate systems include cylindrical coordinates and
spherical coordinates.

Equations of curves

In analytic geometry, any equation involving the coordinates specifies a subset of the plane,
namely the solution set for the equation. For example, the equation y = x corresponds to the set of
all the points on the plane whose x-coordinate and y-coordinate are equal. These points form a
line, and y = x is said to be the equation for this line. In general, linear equations involving x and
y specify lines, quadratic equations specify conic sections, and more complicated equations
describe more complicated figures.

Usually, a single equation corresponds to a curve on the plane. This is not always the case: the
trivial equation x = x specifies the entire plane, and the equation x2 + y2 = 0 specifies only the
single point (0, 0). In three dimensions, a single equation usually gives a surface, and a curve
must be specified as the intersection of two surfaces (see below), or as a system of parametric
equations. The equation x2 + y2 = r2 is the equation for any circle with a radius of r.

The distance formula on the plane follows from the Pythagorean theorem.

Distance and angle

In analytic geometry, geometric notions such as distance and angle measure are defined using
formulas. These definitions are designed to be consistent with the underlying Euclidean
geometry. For example, using Cartesian coordinates on the plane, the distance between two
points (x1, y1) and (x2, y2) is defined by the formula

which can be viewed as a version of the Pythagorean theorem. Similarly, the angle that a line
makes with the horizontal can be defined by the formula

where m is the slope of the line.

Section of a line

This section requires expansion. (November 2012)


This section does not cite any references or sources. Please help improve this section
by adding citations to reliable sources. Unsourced material may be challenged and
removed. (September 2011)

In Analytical Geometry a section of a line can be given by the formula where (c,d)&(e,f) are the
endpoints of the line & m:n is the ratio of division

S(a,b)=(nc+me/m+n, nd+mf/m+n)

Transformations

This section requires expansion. (December 2009)

Transformations are applied to parent functions to turn it into a new function with similar
characteristics. For example, the parent function y=1/x has a horizontal and a vertical asymptote,
and occupies the first and third quadrant, and all of its transformed forms have one horizontal
and vertical asymptote,and occupies either the 1st and 3rd or 2nd and 4th quadrant. In general, if
y = f(x), then it can be transformed into y = af(b(x k)) + h. In the new transformed function, a is
the factor that vertically stretches the function if it is greater than 1 or vertically compresses the
function if it is less than 1, and for negative a values, the function is reflected in the x-axis. The b
value compresses the graph of the function horizontally if greater than 1 and stretches the
function horizontally if less than 1, and like a, reflects the function in the y-axis when it is
negative. The k and h values introduce translations, h, vertical, and k horizontal. Positive h and k
values mean the function is translated to the positive end of its axis and negative meaning
translation towards the negative end.

Transformations can be applied to any geometric equation whether or not the equation represents
a function. Transformations can be considered as individual transactions or in combinations.

Suppose that R(x,y) is a relation in the xy plane. For example


x2 + y2 -1= 0

is the relation that describes the unit circle. The graph of R(x,y) is changed by standard
transformations as follows:

Changing x to x-h moves the graph to the right h units.

Changing y to y-k moves the graph up k units.

Changing x to x/b stretches the graph horizontally by a factor of b. (think of the x as being
dilated)

Changing y to y/a stretches the graph vertically.

Changing x to xcosA+ ysinA and changing y to -xsinA + ycosA rotates the graph by an angle A.

There are other standard transformation not typically studied in elementary analytic geometry
because the transformations change the shape of objects in ways not usually considered. Skewing
is an example of a transformation not usually considered. For more information, consult the
Wikipedia article on affine transformations.

Intersections

While this discussion is limited to the xy-plane, it can easily be extended to higher dimensions.
For two geometric objects P and Q represented by the relations P(x,y) and Q(x,y) the intersection
is the collection of all points (x,y) which are in both relations. For example, P might be the circle
with radius 1 and center (0,0): P = {(x,y) | x2+y2=1} and Q might be the circle with radius 1 and
center (1,0): Q = {(x,y) | (x-1)2+y2=1}. The intersection of these two circles is the collection of
points which make both equations true. Does the point (0,0) make both equations true? Using
(0,0) for (x,y), the equation for Q becomes (0-1)2+02=1 or (-1)2=1 which is true, so (0,0) is in the
relation Q. On the other hand, still using (0,0) for (x,y) the equation for P becomes (0)2+02=1 or
0=1 which is false. (0,0) is not in P so it is not in the intersection.

The intersection of P and Q can be found by solving the simultaneous equations:

x2+y2 = 1

(x-1)2+y2 = 1

Traditional methods include substitution and elimination.

Substitution: Solve the first equation for y in terms of x and then substitute the expression for y
into the second equation.

x2+y2 = 1
y2=1-x2 We then substitute this value for y2 into the other equation:

(x-1)2+(1-x2)=1 and proceed to solve for x:

x2 -2x +1 +1 -x2 =1

-2x = -1

x=

We next place this value of x in either of the original equations and solve for y:

2+y2 = 1

y2 =

So that our intersection has two points:

Elimination: Add (or subtract) a multiple of one equation to the other equation so that one of the
variables is eliminated. For our current example, If we subtract the first equation from the second
we get: (x-1)2-x2=0 The y2 in the first equation is subtracted from the y2 in the second equation
leaving no y term. y has been eliminated. We then solve the remaining equation for x, in the
same way as in the substitution method. x2 -2x +1 +1 -x2 =1 -2x = -1 x= We next place this
value of x in either of the original equations and solve for y: 2+y2 = 1

y2 =

So that our intersection has two points:

For conic sections, as many as 4 points might be in the intersection.


Intercepts

One type of intersection which is widely studied is the intersection of a geometric object with the
x and y coordinate axes.

The intersection of a geometric object and the y-axis is called the y-intercept of the object. The
intersection of a geometric object and the x-axis is called the x-intercept of the object.

For the line y=mx+b, the parameter b specifies the point where the line crosses the y axis.
Depending on the context, either b or the point (0,b) is called the y-intercept.

Themes
Important themes of analytical geometry are

vector space

definition of the plane

distance problems

the dot product, to get the angle of two vectors

the cross product, to get a perpendicular vector of two known vectors (and also their
spatial volume)

intersection problems

conic sections depending on the class, this may include rotation of coordinates and the
general quadratic problems

Ax2 + Bxy + Cy2 +Dx + Ey + F = 0. If the Bxy term is considered, rotations are generally
used.

Many of these problems involve linear algebra.

Example
Here an example of a problem from the United States of America Mathematical Talent Search
that can be solved via analytic geometry:

Problem: In a convex pentagon , the sides have lengths , , , , and , though not
necessarily in that order. Let , , , and be the midpoints of the sides , , , and
, respectively. Let be the midpoint of segment , and be the midpoint of segment
. The length of segment is an integer. Find all possible values for the length of side .

Solution: Without loss of generality, let , , , , and be located at ,


, , , and .

Using the midpoint formula, the points , , , , , and are located at

, , , ,

, and

Using the distance formula,

and

Since has to be an integer,

(see modular arithmetic) so .

Modern analytic geometry


Main article: Algebraic geometry
Further information: Complex geometry

An analytic variety is defined locally as the set of common solutions of several equations
involving analytic functions. It is analogous to the included concept of real or complex algebraic
variety. Any complex manifold is an analytic variety. Since analytic varieties may have singular
points, not all analytic varieties are manifolds.

Analytic geometry is essentially equivalent to real and complex Algebraic geometry, as has been
shown by Jean-Pierre Serre in his paper GAGA, the name of which is French for Algebraic
geometry and analytic geometry. Nevertheless, the two fields remain distinct, as the methods of
proof are quite different and algebraic geometry includes also geometry in finite characteristic.
Notes
1. ^ Boyer, Carl B. (1991). "The Age of Plato and Aristotle". A History of
Mathematics (Second Edition ed.). John Wiley & Sons, Inc.. pp. 9495. ISBN 0-471-
54397-7. "Menaechmus apparently derived these properties of the conic sections and
others as well. Since this material has a strong resemblance to the use of coordinates, as
illustrated above, it has sometimes been maintained that Menaechmus had analytic
geometry. Such a judgment is warranted only in part, for certainly Menaechmus was
unaware that any equation in two unknown quantities determines a curve. In fact, the
general concept of an equation in unknown quantities was alien to Greek thought. It was
shortcomings in algebraic notations that, more than anything else, operated against the
Greek achievement of a full-fledged coordinate geometry."

2. ^ Boyer, Carl B. (1991). "Apollonius of Perga". A History of Mathematics


(Second Edition ed.). John Wiley & Sons, Inc.. pp. 142. ISBN 0-471-54397-7. "The
Apollonian treatise On Determinate Section dealt with what might be called an analytic
geometry of one dimension. It considered the following general problem, using the
typical Greek algebraic analysis in geometric form: Given four points A, B, C, D on a
straight line, determine a fifth point P on it such that the rectangle on AP and CP is in a
given ratio to the rectangle on BP and DP. Here, too, the problem reduces easily to the
solution of a quadratic; and, as in other cases, Apollonius treated the question
exhaustively, including the limits of possibility and the number of solutions."

3. ^ Boyer, Carl B. (1991). "Apollonius of Perga". A History of Mathematics


(Second Edition ed.). John Wiley & Sons, Inc.. pp. 156. ISBN 0-471-54397-7. "The
method of Apollonius in the Conics in many respects are so similar to the modern
approach that his work sometimes is judged to be an analytic geometry anticipating that
of Descartes by 1800 years. The application of references lines in general, and of a
diameter and a tangent at its extremity in particular, is, of course, not essentially different
from the use fo a coordinate frame, whether rectangular or, more generally, oblique.
Distances measured along the diameter from the point of tangency are the abscissas, and
segments parallel to the tangent and intercepted between the axis and the curve are the
ordinates. The Apollonian relationship between these abscissas and the corresponding
ordinates are nothing more nor less than rhetorical forms of the equations of the curves.
However, Greek geometric algebra did not provide for negative magnitudes; moreover,
the coordinate system was in every case superimposed a posteriori upon a given curve in
order to study its properties. There appear to be no cases in ancient geometry in which a
coordinate frame of reference was laid down a priori for purposes of graphical
representation of an equation or relationship, whether symbolically or rhetorically
expressed. Of Greek geometry we may say that equations are determined by curves, but
not that curves are determined by equations. Coordinates, variables, and equations were
subsidiary notions derived from a specific geometric situation; [...] That Apollonius, the
greatest geometer of antiquity, failed to develop analytic geometry, was probably the
result of a poverty of curves rather than of thought. General methods are not necessary
when problems concern always one of a limited number of particular cases."
4. ^ a b c Boyer (1991). "The Arabic Hegemony". pp. 241242. "Omar Khayyam (ca.
10501123), the "tent-maker," wrote an Algebra that went beyond that of al-Khwarizmi
to include equations of third degree. Like his Arab predecessors, Omar Khayyam
provided for quadratic equations both arithmetic and geometric solutions; for general
cubic equations, he believed (mistakenly, as the sixteenth century later showed),
arithmetic solutions were impossible; hence he gave only geometric solutions. The
scheme of using intersecting conics to solve cubics had been used earlier by
Menaechmus, Archimedes, and Alhazan, but Omar Khayyam took the praiseworthy step
of generalizing the method to cover all third-degree equations (having positive roots). ..
For equations of higher degree than three, Omar Khayyam evidently did not envision
similar geometric methods, for space does not contain more than three dimensions, ...
One of the most fruitful contributions of Arabic eclecticism was the tendency to close the
gap between numerical and geometric algebra. The decisive step in this direction came
much later with Descartes, but Omar Khayyam was moving in this direction when he
wrote, "Whoever thinks algebra is a trick in obtaining unknowns has thought it in vain.
No attention should be paid to the fact that algebra and geometry are different in
appearance. Algebras are geometric facts which are proved.""

5. ^ Glen M. Cooper (2003). "Omar Khayyam, the Mathmetician", The Journal of


the American Oriental Society 123.

6. ^ Stillwell, John (2004). "Analytic Geometry". Mathematics and its History


(Second Edition ed.). Springer Science + Business Media Inc.. pp. 105. ISBN 0-387-
95336-1. "the two founders of analytic geometry, Fermat and Descartes, were both
strongly influenced by these developments."

7. ^ Cooke, Roger (1997). "The Calculus". The History of Mathematics: A Brief


Course. Wiley-Interscience. pp. 326. ISBN 0-471-18082-3. "The person who is popularly
credited with being the discoverer of analytic geometry was the philosopher Ren
Descartes (15961650), one of the most influential thinkers of the modern era."

8. ^ a b Katz 1998, pg. 442

9. ^ Katz 1998, pg. 436


1
History of geometry

From Wikipedia, the free encyclopedia

Jump to: navigation, search

1
Table of Geometry, from the 1728 Cyclopaedia.

Geometry

Oxyrhynchus papyrus (P.Oxy. I 29) showing


fragment of Euclid's Elements

History of geometry

Branches[show]

Research areas[show]

Important concepts[show]

Geometers[show]
v

History of science

Background[hide]

Theories/sociology

Historiography

Pseudoscience

By era[hide]

In early cultures

in Classical Antiquity
In the Middle Ages

In the Renaissance

Scientific revolution

Romanticism in science

By culture[hide]

African

Byzantine

Chinese

Indian

Islamic

Natural sciences[hide]

Astronomy

Biology

Botany

Chemistry

Ecology

Evolution

Geology

Geophysics

Paleontology
Physics

Mathematics[hide]

Algebra

Calculus

Combinatorics

Geometry

Logic

Probability

Statistics

Trigonometry

Social sciences[hide]

Anthropology

Economics

Geography

Linguistics

Political science

Psychology

Sociology

Sustainability

Technology[hide]
Agricultural science

Computer science

Materials science

Medicine[hide]

Medicine

Navigational pages[hide]

Timelines

Portal

Categories

Geometry (Greek ; geo = earth, metria = measure) arose as the field of knowledge
dealing with spatial relationships. Geometry was one of the two fields of pre-modern
mathematics, the other being the study of numbers (arithmetic).

Classic geometry was focused in compass and straightedge constructions. Geometry was
revolutionized by Euclid, who introduced mathematical rigor and the axiomatic method still in
use today. His book, The Elements is widely considered the most influential textbook of all time,
and was known to all educated people in the West until the middle of the 20th century.[1]

In modern times, geometric concepts have been generalized to a high level of abstraction and
complexity, and have been subjected to the methods of calculus and abstract algebra, so that
many modern branches of the field are barely recognizable as the descendants of early geometry.
(See areas of mathematics and algebraic geometry.)
Contents

1 Early geometry

o 1.1 Egyptian geometry

o 1.2 Babylonian geometry

2 Greek geometry

o 2.1 Classical Greek geometry

2.1.1 Thales and Pythagoras

2.1.2 Plato

o 2.2 Hellenistic geometry

2.2.1 Euclid

2.2.2 Archimedes

2.2.3 After Archimedes

3 Indian geometry

o 3.1 Vedic period

o 3.2 Classical period

4 Chinese geometry

o 4.1 The Nine Chapters on the Mathematical Art

5 Islamic geometry

o 5.1 Thabit family and other early geometers

o 5.2 Geometric architecture

6 Modern geometry

o 6.1 The 17th century

o 6.2 The 18th and 19th centuries


6.2.1 Non-Euclidean geometry

6.2.2 Introduction of mathematical rigor

6.2.3 Analysis situs, or topology

o 6.3 The 20th century

7 Timeline

8 See also

9 Notes

10 References

11 External links

Early geometry

The earliest recorded beginnings of geometry can be traced to early peoples, who discovered
obtuse triangles in the ancient Indus Valley (see Harappan Mathematics), and ancient Babylonia
(see Babylonian mathematics) from around 3000 BC. Early geometry was a collection of
empirically discovered principles concerning lengths, angles, areas, and volumes, which were
developed to meet some practical need in surveying, construction, astronomy, and various crafts.
Among these were some surprisingly sophisticated principles, and a modern mathematician
might be hard put to derive some of them without the use of calculus. For example, both the
Egyptians and the Babylonians were aware of versions of the Pythagorean theorem about 1500
years before Pythagoras; the Egyptians had a correct formula for the volume of a frustum of a
square pyramid;

Egyptian geometry

Main article: Egyptian mathematics

The ancient Egyptians knew that they could approximate the area of a circle as follows:[2]

Area of Circle [ (Diameter) x 8/9 ]2.[2]

Problem 30 of the Ahmes papyrus uses these methods to calculate the area of a circle, according
to a rule that the area is equal to the square of 8/9 of the circle's diameter. This assumes that is
4(8/9) (or 3.160493...), with an error of slightly over 0.63 percent. This value was slightly less
accurate than the calculations of the Babylonians (25/8 = 3.125, within 0.53 percent), but was not
otherwise surpassed until Archimedes' approximation of 211875/67441 = 3.14163, which had an
error of just over 1 in 10,000.

Interestingly, Ahmes knew of the modern 22/7 as an approximation for pi, and used it to split a
hekat, hekat x 22/x x 7/22 = hekat; however, Ahmes continued to use the traditional 256/81 value
for pi for computing his hekat volume found in a cylinder.

Problem 48 involved using a square with side 9 units. This square was cut into a 3x3 grid. The
diagonal of the corner squares were used to make an irregular octagon with an area of 63 units.
This gave a second value for of 3.111...

The two problems together indicate a range of values for Pi between 3.11 and 3.16.

Problem 14 in the Moscow Mathematical Papyrus gives the only ancient example finding the
volume of a frustum of a pyramid, describing the correct formula:

Babylonian geometry

Main article: Babylonian mathematics

The Babylonians may have known the general rules for measuring areas and volumes. They
measured the circumference of a circle as three times the diameter and the area as one-twelfth the
square of the circumference, which would be correct if is estimated as 3. The volume of a
cylinder was taken as the product of the base and the height, however, the volume of the frustum
of a cone or a square pyramid was incorrectly taken as the product of the height and half the sum
of the bases. The Pythagorean theorem was also known to the Babylonians. Also, there was a
recent discovery in which a tablet used as 3 and 1/8. The Babylonians are also known for the
Babylonian mile, which was a measure of distance equal to about seven miles today. This
measurement for distances eventually was converted to a time-mile used for measuring the travel
of the Sun, therefore, representing time.[3]

Greek geometry

See also: Greek mathematics

Classical Greek geometry

For the ancient Greek mathematicians, geometry was the crown jewel of their sciences, reaching
a completeness and perfection of methodology that no other branch of their knowledge had
attained. They expanded the range of geometry to many new kinds of figures, curves, surfaces,
and solids; they changed its methodology from trial-and-error to logical deduction; they
recognized that geometry studies "eternal forms", or abstractions, of which physical objects are
only approximations; and they developed the idea of the "axiomatic method", still in use today.

Thales and Pythagoras

Pythagorean theorem: a2 + b2 = c2

Thales (635-543 BC) of Miletus (now in southwestern Turkey), was the first to whom deduction
in mathematics is attributed. There are five geometric propositions for which he wrote deductive
proofs, though his proofs have not survived. Pythagoras (582-496 BC) of Ionia, and later, Italy,
then colonized by Greeks, may have been a student of Thales, and traveled to Babylon and
Egypt. The theorem that bears his name may not have been his discovery, but he was probably
one of the first to give a deductive proof of it. He gathered a group of students around him to
study mathematics, music, and philosophy, and together they discovered most of what high
school students learn today in their geometry courses. In addition, they made the profound
discovery of incommensurable lengths and irrational numbers.

Plato
Plato (427-347 BC), the philosopher most esteemed by the Greeks, had inscribed above the
entrance to his famous school, "Let none ignorant of geometry enter here." Though he was not a
mathematician himself, his views on mathematics had great influence. Mathematicians thus
accepted his belief that geometry should use no tools but compass and straightedge never
measuring instruments such as a marked ruler or a protractor, because these were a workmans
tools, not worthy of a scholar. This dictum led to a deep study of possible compass and
straightedge constructions, and three classic construction problems: how to use these tools to
trisect an angle, to construct a cube twice the volume of a given cube, and to construct a square
equal in area to a given circle. The proofs of the impossibility of these constructions, finally
achieved in the 19th century, led to important principles regarding the deep structure of the real
number system. Aristotle (384-322 BC), Platos greatest pupil, wrote a treatise on methods of
reasoning used in deductive proofs (see Logic) which was not substantially improved upon until
the 19th century.
Hellenistic geometry

Euclid

Statue of Euclid in the Oxford University Museum of Natural History.

Woman teaching geometry. Illustration at the beginning of a medieval translation of


Euclid's Elements, (c. 1310)
Euclid (c. 325-265 BC), of Alexandria, probably a student of one of Platos students, wrote a
treatise in 13 books (chapters), titled The Elements of Geometry, in which he presented geometry
in an ideal axiomatic form, which came to be known as Euclidean geometry. The treatise is not a
compendium of all that the Hellenistic mathematicians knew at the time about geometry; Euclid
himself wrote eight more advanced books on geometry. We know from other references that
Euclids was not the first elementary geometry textbook, but it was so much superior that the
others fell into disuse and were lost. He was brought to the university at Alexandria by Ptolemy
I, King of Egypt.

The Elements began with definitions of terms, fundamental geometric principles (called axioms
or postulates), and general quantitative principles (called common notions) from which all the
rest of geometry could be logically deduced. Following are his five axioms, somewhat
paraphrased to make the English easier to read.

1. Any two points can be joined by a straight line.

2. Any finite straight line can be extended in a straight line.

3. A circle can be drawn with any center and any radius.

4. All right angles are equal to each other.

5. If two straight lines in a plane are crossed by another straight line (called the
transversal), and the interior angles between the two lines and the
transversal lying on one side of the transversal add up to less than two right
angles, then on that side of the transversal, the two lines extended will
intersect (also called the parallel postulate).

Archimedes
Archimedes (287-212 BC), of Syracuse, Sicily, when it was a Greek city-state, is often
considered to be the greatest of the Greek mathematicians, and occasionally even named as one
of the three greatest of all time (along with Isaac Newton and Carl Friedrich Gauss). Had he not
been a mathematician, he would still be remembered as a great physicist, engineer, and inventor.
In his mathematics, he developed methods very similar to the coordinate systems of analytic
geometry, and the limiting process of integral calculus. The only element lacking for the creation
of these fields was an efficient algebraic notation in which to express his concepts[citation needed].
After Archimedes

Geometry was connected to the divine for most medieval scholars. The compass in
this 13th-century manuscript is a symbol of God's act of Creation.

After Archimedes, Hellenistic mathematics began to decline. There were a few minor stars yet to
come, but the golden age of geometry was over. Proclus (410-485), author of Commentary on
the First Book of Euclid, was one of the last important players in Hellenistic geometry. He was a
competent geometer, but more importantly, he was a superb commentator on the works that
preceded him. Much of that work did not survive to modern times, and is known to us only
through his commentary. The Roman Republic and Empire that succeeded and absorbed the
Greek city-states produced excellent engineers, but no mathematicians of note.

The great Library of Alexandria was later burned. There is a growing consensus among
historians that the Library of Alexandria likely suffered from several destructive events, but that
the destruction of Alexandria's pagan temples in the late 4th century was probably the most
severe and final one. The evidence for that destruction is the most definitive and secure. Caesar's
invasion may well have led to the loss of some 40,000-70,000 scrolls in a warehouse adjacent to
the port (as Luciano Canfora argues, they were likely copies produced by the Library intended
for export), but it is unlikely to have affected the Library or Museum, given that there is ample
evidence that both existed later.

Civil wars, decreasing investments in maintenance and acquisition of new scrolls and generally
declining interest in non-religious pursuits likely contributed to a reduction in the body of
material available in the Library, especially in the 4th century. The Serapeum was certainly
destroyed by Theophilus in 391, and the Museum and Library may have fallen victim to the same
campaign.
Indian geometry

See also: Indian mathematics

Vedic period

Rigveda manuscript in Devanagari.

The Satapatha Brahmana (9th century BCE) contains rules for ritual geometric constructions
that are similar to the Sulba Sutras.[4]

The ulba Stras (literally, "Aphorisms of the Chords" in Vedic Sanskrit) (c. 700-400 BCE) list
rules for the construction of sacrificial fire altars.[5] Most mathematical problems considered in
the ulba Stras spring from "a single theological requirement,"[6] that of constructing fire altars
which have different shapes but occupy the same area. The altars were required to be constructed
of five layers of burnt brick, with the further condition that each layer consist of 200 bricks and
that no two adjacent layers have congruent arrangements of bricks.[6]

According to (Hayashi 2005, p. 363), the ulba Stras contain "the earliest extant verbal
expression of the Pythagorean Theorem in the world, although it had already been known to the
Old Babylonians."

The diagonal rope (akay-rajju) of an oblong (rectangle) produces both which the flank
(prvamni) and the horizontal (tiryamn) <ropes> produce separately."[7]

Since the statement is a stra, it is necessarily compressed and what the ropes produce is not
elaborated on, but the context clearly implies the square areas constructed on their lengths, and
would have been explained so by the teacher to the student.[7]
They contain lists of Pythagorean triples,[8] which are particular cases of Diophantine equations.
[9]
They also contain statements (that with hindsight we know to be approximate) about squaring
the circle and "circling the square."[10]

Baudhayana (c. 8th century BCE) composed the Baudhayana Sulba Sutra, the best-known Sulba
Sutra, which contains examples of simple Pythagorean triples, such as: , ,
[11]
, , and as well as a statement of the Pythagorean
theorem for the sides of a square: "The rope which is stretched across the diagonal of a square
produces an area double the size of the original square."[11] It also contains the general statement
of the Pythagorean theorem (for the sides of a rectangle): "The rope stretched along the length of
the diagonal of a rectangle makes an area which the vertical and horizontal sides make
together."[11]

According to mathematician S. G. Dani, the Babylonian cuneiform tablet Plimpton 322 written c.
1850 BCE[12] "contains fifteen Pythagorean triples with quite large entries, including (13500,
12709, 18541) which is a primitive triple,[13] indicating, in particular, that there was sophisticated
understanding on the topic" in Mesopotamia in 1850 BCE. "Since these tablets predate the
Sulbasutras period by several centuries, taking into account the contextual appearance of some of
the triples, it is reasonable to expect that similar understanding would have been there in
India."[14] Dani goes on to say:

"As the main objective of the Sulvasutras was to describe the constructions of altars and the
geometric principles involved in them, the subject of Pythagorean triples, even if it had been well
understood may still not have featured in the Sulvasutras. The occurrence of the triples in the
Sulvasutras is comparable to mathematics that one may encounter in an introductory book on
architecture or another similar applied area, and would not correspond directly to the overall
knowledge on the topic at that time. Since, unfortunately, no other contemporaneous sources
have been found it may never be possible to settle this issue satisfactorily."[14]

In all, three Sulba Sutras were composed. The remaining two, the Manava Sulba Sutra composed
by Manava (fl. 750-650 BCE) and the Apastamba Sulba Sutra, composed by Apastamba (c. 600
BCE), contained results similar to the Baudhayana Sulba Sutra.

Classical period

In the Bakhshali manuscript, there is a handful of geometric problems (including problems about
volumes of irregular solids). The Bakhshali manuscript also "employs a decimal place value
system with a dot for zero."[15] Aryabhata's Aryabhatiya (499 CE) includes the computation of
areas and volumes.
Brahmagupta wrote his astronomical work Brhma Sphua Siddhnta in 628 CE. Chapter 12,
containing 66 Sanskrit verses, was divided into two sections: "basic operations" (including cube
roots, fractions, ratio and proportion, and barter) and "practical mathematics" (including mixture,
mathematical series, plane figures, stacking bricks, sawing of timber, and piling of grain).[16] In
the latter section, he stated his famous theorem on the diagonals of a cyclic quadrilateral:[16]

Brahmagupta's theorem: If a cyclic quadrilateral has diagonals that are perpendicular to each
other, then the perpendicular line drawn from the point of intersection of the diagonals to any
side of the quadrilateral always bisects the opposite side.

Chapter 12 also included a formula for the area of a cyclic quadrilateral (a generalization of
Heron's formula), as well as a complete description of rational triangles (i.e. triangles with
rational sides and rational areas).

Brahmagupta's formula: The area, A, of a cyclic quadrilateral with sides of lengths a, b, c, d,


respectively, is given by

where s, the semiperimeter, given by:

Brahmagupta's Theorem on rational triangles: A triangle with rational sides and


rational area is of the form:

for some rational numbers and .[17]

Chinese geometry

See also: Chinese mathematics


The Nine Chapters on the Mathematical Art, first compiled in 179 AD, with added
commentary in the 3rd century by Liu Hui.

The Sea Island Mathematical Manual, Liu Hui, 3rd century.

The first definitive work (or at least oldest existent) on geometry in China was the Mo Jing, the
Mohist canon of the early utilitarian philosopher Mozi (470-390 BC). It was compiled years after
his death by his later followers around the year 330 BC.[18] Although the Mo Jing is the oldest
existent book on geometry in China, there is the possibility that even older written material
exists. However, due to the infamous Burning of the Books in the political maneauver by the Qin
Dynasty ruler Qin Shihuang (r. 221-210 BC), multitudes of written literature created before his
time was purged. In addition, the Mo Jing presents geometrical concepts in mathematics that are
perhaps too advanced not to have had a previous geometrical base or mathematic background to
work upon.

The Mo Jing described various aspects of many fields associated with physical science, and
provided a small wealth of information on mathematics as well. It provided an 'atomic' definition
of the geometric point, stating that a line is separated into parts, and the part which has no
remaining parts (i.e. cannot be divided into smaller parts) and thus forms the extreme end of a
line is a point.[18] Much like Euclid's first and third definitions and Plato's 'beginning of a line',
the Mo Jing stated that "a point may stand at the end (of a line) or at its beginning like a head-
presentation in childbirth. (As to its invisibility) there is nothing similar to it."[19] Similar to the
atomists of Democritus, the Mo Jing stated that a point is the smallest unit, and cannot be cut in
half, since 'nothing' cannot be halved.[19] It stated that two lines of equal length will always finish
at the same place,[19] while providing definitions for the comparison of lengths and for parallels,
[20]
along with principles of space and bounded space.[21] It also described the fact that planes
without the quality of thickness cannot be piled up since they cannot mutually touch.[22] The book
provided definitions for circumference, diameter, and radius, along with the definition of
volume.[23]

The Han Dynasty (202 BC-220 AD) period of China witnessed a new flourishing of
mathematics. One of the oldest Chinese mathematical texts to present geometric progressions
was the Sun sh sh of 186 BC, during the Western Han era. The mathematician, inventor, and
astronomer Zhang Heng (78-139 AD) used geometrical formulas to solve mathematical
problems. Although rough estimates for pi () were given in the Zhou Li (compiled in the 2nd
century BC),[24] it was Zhang Heng who was the first to make a concerted effort at creating a
more accurate formula for pi. This in turn would be made more accurate by later Chinese such as
Zu Chongzhi (429-500 AD). Zhang Heng approximated pi as 730/232 (or approx 3.1466),
although he used another formula of pi in finding a spherical volume, using the square root of 10
(or approx 3.162) instead. Zu Chongzhi's best approximation was between 3.1415926 and
3.1415927, with 355113 (, Mil, detailed approximation) and 227 (, Yuel, rough
approximation) being the other notable approximation.[25] In comparison to later works, the
formula for pi given by the French mathematician Franciscus Vieta (1540-1603) fell halfway
between Zu's approximations.

The Nine Chapters on the Mathematical Art

The Nine Chapters on the Mathematical Art, the title of which first appeared by 179 AD on a
bronze inscription, was edited and commented on by the 3rd century mathematician Liu Hui
from the Kingdom of Cao Wei. This book included many problems where geometry was applied,
such as finding surface areas for squares and circles, the volumes of solids in various three
dimensional shapes, and included the use of the Pythagorean theorem. The book provided
illustrated proof for the Pythagorean theorem,[26] contained a written dialogue between of the
earlier Duke of Zhou and Shang Gao on the properties of the right angle triangle and the
Pythagorean theorem, while also referring to the astronomical gnomon, the circle and square, as
well as measurements of heights and distances.[27] The editor Liu Hui listed pi as 3.141014 by
using a 192 sided polygon, and then calculated pi as 3.14159 using a 3072 sided polygon. This
was more accurate than Liu Hui's contemporary Wang Fan, a mathematician and astronomer
from Eastern Wu, would render pi as 3.1555 by using 14245.[28] Liu Hui also wrote of mathematical
surveying to calculate distance measurements of depth, height, width, and surface area. In terms
of solid geometry, he figured out that a wedge with rectangular base and both sides sloping could
be broken down into a pyramid and a tetrahedral wedge.[29] He also figured out that a wedge with
trapezoid base and both sides sloping could be made to give two tetrahedral wedges separated by
a pyramid.[29] Furthermore, Liu Hui described Cavalieri's principle on volume, as well as
Gaussian elimination. From the Nine Chapters, it listed the following geometrical formulas that
were known by the time of the Former Han Dynasty (202 BCE9 CE).

Areas for the[30]

Square Rhomboid

Rectangle Trapezoid

Circle Double
trapezium
Isosceles
triangle Segment of a
circle

Annulus ('ring'
between two
concentric
circles)

Volumes for the[29]

Parallelepiped Cube Frustum of a


with two wedge of the
square Prism second type
surfaces (used for
Wedge with applications in
Parallelepiped rectangular engineering)
with no square base and both
surfaces sides sloping Cylinder

Pyramid Wedge with Cone with


trapezoid circular base
Frustum of base and both
pyramid with sides sloping Frustum of a
square base cone
Tetrahedral
Frustum of wedge Sphere
pyramid with
rectangular
base of
unequal sides

Continuing the geometrical legacy of ancient China, there were many later figures to come,
including the famed astronomer and mathematician Shen Kuo (1031-1095 CE), Yang Hui (1238-
1298) who discovered Pascal's Triangle, Xu Guangqi (1562-1633), and many others.

Islamic geometry

See also: Islamic mathematics

Page from the Al-Jabr wa-al-Muqabilah

The Islamic Caliphate established across the Middle East, North Africa, Spain, Portugal, Persia
and parts of Persia, began around 640 CE. Islamic mathematics during this period was primarily
algebraic rather than geometric, though there were important works on geometry. Scholarship in
Europe declined and eventually the Hellenistic works of antiquity were lost to them, and
survived only in the Islamic centers of learning.

Although the Muslim mathematicians are most famed for their work on algebra, number theory
and number systems, they also made considerable contributions to geometry, trigonometry and
mathematical astronomy, and were responsible for the development of algebraic geometry.
Geometrical magnitudes were treated as "algebraic objects" by most Muslim mathematicians
however.

The successors of Muammad ibn Ms al-wrizm who was Persian Scholar, mathematician
and Astronomer who invented the Algorithm in Mathematics which is the base for Computer
Science (born 780) undertook a systematic application of arithmetic to algebra, algebra to
arithmetic, both to trigonometry, algebra to the Euclidean theory of numbers, algebra to
geometry, and geometry to algebra. This was how the creation of polynomial algebra,
combinatorial analysis, numerical analysis, the numerical solution of equations, the new
elementary theory of numbers, and the geometric construction of equations arose.

Al-Mahani (born 820) conceived the idea of reducing geometrical problems such as duplicating
the cube to problems in algebra. Al-Karaji (born 953) completely freed algebra from geometrical
operations and replaced them with the arithmetical type of operations which are at the core of
algebra today.

An engraving by Albrecht Drer featuring Mashallah, from the title page of the De
scientia motus orbis (Latin version with engraving, 1504). As in many medieval
illustrations, the compass here is an icon of religion as well as science, in reference
to God as the architect of creation

Thabit family and other early geometers

Thabit ibn Qurra (known as Thebit in Latin) (born 836) contributed to a number of areas in
mathematics, where he played an important role in preparing the way for such important
mathematical discoveries as the extension of the concept of number to (positive) real numbers,
integral calculus, theorems in spherical trigonometry, analytic geometry, and non-Euclidean
geometry. In astronomy Thabit was one of the first reformers of the Ptolemaic system, and in
mechanics he was a founder of statics. An important geometrical aspect of Thabit's work was his
book on the composition of ratios. In this book, Thabit deals with arithmetical operations applied
to ratios of geometrical quantities. The Greeks had dealt with geometric quantities but had not
thought of them in the same way as numbers to which the usual rules of arithmetic could be
applied. By introducing arithmetical operations on quantities previously regarded as geometric
and non-numerical, Thabit started a trend which led eventually to the generalisation of the
number concept.

In some respects, Thabit is critical of the ideas of Plato and Aristotle, particularly regarding
motion. It would seem that here his ideas are based on an acceptance of using arguments
concerning motion in his geometrical arguments. Another important contribution Thabit made to
geometry was his generalization of the Pythagorean theorem, which he extended from special
right triangles to all triangles in general, along with a general proof.[31]

Ibrahim ibn Sinan ibn Thabit (born 908), who introduced a method of integration more general
than that of Archimedes, and al-Quhi (born 940) were leading figures in a revival and
continuation of Greek higher geometry in the Islamic world. These mathematicians, and in
particular Ibn al-Haytham, studied optics and investigated the optical properties of mirrors made
from conic sections.

Astronomy, time-keeping and geography provided other motivations for geometrical and
trigonometrical research. For example Ibrahim ibn Sinan and his grandfather Thabit ibn Qurra
both studied curves required in the construction of sundials. Abu'l-Wafa and Abu Nasr Mansur
both applied spherical geometry to astronomy.

Geometric architecture

Recent discoveries have shown that geometrical quasicrystal patterns were first employed in the
girih tiles found in medieval Islamic architecture dating back over five centuries ago. In 2007,
Professor Peter Lu of Harvard University and Professor Paul Steinhardt of Princeton University
published a paper in the journal Science suggesting that girih tilings possessed properties
consistent with self-similar fractal quasicrystalline tilings such as the Penrose tilings, predating
them by five centuries.[32][33]

Modern geometry

The 17th century

When Europe began to emerge from its Dark Ages, the Hellenistic and Islamic texts on geometry
found in Islamic libraries were translated from Arabic into Latin. The rigorous deductive
methods of geometry found in Euclids Elements of Geometry were relearned, and further
development of geometry in the styles of both Euclid (Euclidean geometry) and Khayyam
(algebraic geometry) continued, resulting in an abundance of new theorems and concepts, many
of them very profound and elegant.
Discourse on Method by Ren Descartes

In the early 17th century, there were two important developments in geometry. The first and most
important was the creation of analytic geometry, or geometry with coordinates and equations, by
Ren Descartes (15961650) and Pierre de Fermat (16011665). This was a necessary precursor
to the development of calculus and a precise quantitative science of physics. The second
geometric development of this period was the systematic study of projective geometry by Girard
Desargues (15911661). Projective geometry is the study of geometry without measurement, just
the study of how points align with each other. There had been some early work in this area by
Hellenistic geometers, notably Pappus (c. 340). The greatest flowering of the field occurred with
Jean-Victor Poncelet (17881867).

In the late 17th century, calculus was developed independently and almost simultaneously by
Isaac Newton (16421727) and Gottfried Wilhelm Leibniz (16461716). This was the beginning
of a new field of mathematics now called analysis. Though not itself a branch of geometry, it is
applicable to geometry, and it solved two families of problems that had long been almost
intractable: finding tangent lines to odd curves, and finding areas enclosed by those curves. The
methods of calculus reduced these problems mostly to straightforward matters of computation.

The 18th and 19th centuries

Non-Euclidean geometry
The old problem of proving Euclids Fifth Postulate, the "Parallel Postulate", from his first four
postulates had never been forgotten. Beginning not long after Euclid, many attempted
demonstrations were given, but all were later found to be faulty, through allowing into the
reasoning some principle which itself had not been proved from the first four postulates. Though
Omar Khayym was also unsuccessful in proving the parallel postulate, his criticisms of Euclid's
theories of parallels and his proof of properties of figures in non-Euclidean geometries
contributed to the eventual development of non-Euclidean geometry. By 1700 a great deal had
been discovered about what can be proved from the first four, and what the pitfalls were in
attempting to prove the fifth. Saccheri, Lambert, and Legendre each did excellent work on the
problem in the 18th century, but still fell short of success. In the early 19th century, Gauss,
Johann Bolyai, and Lobatchewsky, each independently, took a different approach. Beginning to
suspect that it was impossible to prove the Parallel Postulate, they set out to develop a self-
consistent geometry in which that postulate was false. In this they were successful, thus creating
the first non-Euclidean geometry. By 1854, Bernhard Riemann, a student of Gauss, had applied
methods of calculus in a ground-breaking study of the intrinsic (self-contained) geometry of all
smooth surfaces, and thereby found a different non-Euclidean geometry. This work of Riemann
later became fundamental for Einstein's theory of relativity.

William Blake's "Newton" is a demonstration of his opposition to the 'single-vision' of


scientific materialism; here, Isaac Newton is shown as 'divine geometer' (1795)

It remained to be proved mathematically that the non-Euclidean geometry was just as self-
consistent as Euclidean geometry, and this was first accomplished by Beltrami in 1868. With
this, non-Euclidean geometry was established on an equal mathematical footing with Euclidean
geometry.

While it was now known that different geometric theories were mathematically possible, the
question remained, "Which one of these theories is correct for our physical space?" The
mathematical work revealed that this question must be answered by physical experimentation,
not mathematical reasoning, and uncovered the reason why the experimentation must involve
immense (interstellar, not earth-bound) distances. With the development of relativity theory in
physics, this question became vastly more complicated.
Introduction of mathematical rigor
All the work related to the Parallel Postulate revealed that it was quite difficult for a geometer to
separate his logical reasoning from his intuitive understanding of physical space, and, moreover,
revealed the critical importance of doing so. Careful examination had uncovered some logical
inadequacies in Euclid's reasoning, and some unstated geometric principles to which Euclid
sometimes appealed. This critique paralleled the crisis occurring in calculus and analysis
regarding the meaning of infinite processes such as convergence and continuity. In geometry,
there was a clear need for a new set of axioms, which would be complete, and which in no way
relied on pictures we draw or on our intuition of space. Such axioms were given by David
Hilbert in 1894 in his dissertation Grundlagen der Geometrie (Foundations of Geometry). Some
other complete sets of axioms had been given a few years earlier, but did not match Hilbert's in
economy, elegance, and similarity to Euclid's axioms.

Analysis situs, or topology


In the mid-18th century, it became apparent that certain progressions of mathematical reasoning
recurred when similar ideas were studied on the number line, in two dimensions, and in three
dimensions. Thus the general concept of a metric space was created so that the reasoning could
be done in more generality, and then applied to special cases. This method of studying calculus-
and analysis-related concepts came to be known as analysis situs, and later as topology. The
important topics in this field were properties of more general figures, such as connectedness and
boundaries, rather than properties like straightness, and precise equality of length and angle
measurements, which had been the focus of Euclidean and non-Euclidean geometry. Topology
soon became a separate field of major importance, rather than a sub-field of geometry or
analysis.

The 20th century

Developments in algebraic geometry included the study of curves and surfaces over finite fields
as demonstrated by the works of among others Andr Weil, Alexander Grothendieck, and Jean-
Pierre Serre as well as over the real or complex numbers. Finite geometry itself, the study of
spaces with only finitely many points, found applications in coding theory and cryptography.
With the advent of the computer, new disciplines such as computational geometry or digital
geometry deal with geometric algorithms, discrete representations of geometric data, and so
forth.

Timeline

Main article: Timeline of geometry

See also
Wikisource has original text related to this article:

Flatland

Flatland, a book by "A. Square" about two and three-dimensional space, to


understand the concept of four dimensions

History of mathematics

Important publications in geometry.

Interactive geometry software

List of geometry topics

Notes

1. ^ Howard Eves, An Introduction to the History of Mathematics,


Saunders, 1990, ISBN 0-03-029558-0 p. 141: "No work, except The Bible, has
been more widely used...."

2. ^ a b Ray C. Jurgensen, Alfred J. Donnelly, and Mary P. Dolciani. Editorial


Advisors Andrew M. Gleason, Albert E. Meder, Jr. Modern School Mathematics:
Geometry (Student's Edition). Houghton Mifflin Company, Boston, 1972, p.
52. ISBN 0-395-13102-2. Teachers Edition ISBN 0-395-13103-0.

3. ^ Eves, Chapter 2.

4. ^ A. Seidenberg, 1978. The origin of mathematics. Archive for the


history of Exact Sciences, vol 18.

5. ^ (Staal 1999)

a b
6. ^ (Hayashi 2003, p. 118)

a b
7. ^ (Hayashi 2005, p. 363)

8. ^ Pythagorean triples are triples of integers with the property:


. Thus, , , etc.

9. ^ (Cooke 2005, p. 198): "The arithmetic content of the ulva Stras


consists of rules for finding Pythagorean triples such as (3, 4, 5), (5, 12, 13),
(8, 15, 17), and (12, 35, 37). It is not certain what practical use these
arithmetic rules had. The best conjecture is that they were part of religious
ritual. A Hindu home was required to have three fires burning at three
different altars. The three altars were to be of different shapes, but all three
were to have the same area. These conditions led to certain "Diophantine"
problems, a particular case of which is the generation of Pythagorean triples,
so as to make one square integer equal to the sum of two others."

10. ^ (Cooke 2005, pp. 199200): "The requirement of three altars of equal
areas but different shapes would explain the interest in transformation of
areas. Among other transformation of area problems the Hindus considered in
particular the problem of squaring the circle. The Bodhayana Sutra states the
converse problem of constructing a circle equal to a given square. The
following approximate construction is given as the solution.... this result is
only approximate. The authors, however, made no distinction between the
two results. In terms that we can appreciate, this construction gives a value
for of 18(3 22), which is about 3.088."

a b c
11. ^ (Joseph 2000, p. 229)

12. ^ Mathematics Department, University of British Columbia, The


Babylonian tabled Plimpton 322.

13. ^ Three positive integers form a primitive Pythagorean triple if


and if the highest common factor of is 1. In the particular
Plimpton322 example, this means that and that
the three numbers do not have any common factors. However some scholars
have disputed the Pythagorean interpretation of this tablet; see Plimpton 322
for details.

a b
14. ^ (Dani 2003)

15. ^ (Hayashi 2005, p. 371)

a b
16. ^ (Hayashi 2003, pp. 121122)

17. ^ (Stillwell 2004, p. 77)

a b
18. ^ Needham, Volume 3, 91.

a b c
19. ^ Needham, Volume 3, 92.

20. ^ Needham, Volume 3, 92-93.

21. ^ Needham, Volume 3, 93.

22. ^ Needham, Volume 3, 93-94.

23. ^ Needham, Volume 3, 94.

24. ^ Needham, Volume 3, 99.


25. ^ Needham, Volume 3, 101.

26. ^ Needham, Volume 3, 22.

27. ^ Needham, Volume 3, 21.

28. ^ Needham, Volume 3, 100.

a b c
29. ^ Needham, Volume 3, 9899.

30. ^ Needham, Volume 3, 98.

31. ^ Sayili, Aydin (1960). "Thabit ibn Qurra's Generalization of the


Pythagorean Theorem". Isis 51 (1): 3537.

32. ^ Peter J. Lu and Paul J. Steinhardt (2007), "Decagonal and Quasi-


crystalline Tilings in Medieval Islamic Architecture", Science 315 (5815):
11061110, Bibcode 2007Sci...315.1106L, doi:10.1126/science.1135491,
PMID 17322056.

33. ^ Supplemental figures

Eudoxus of Cnidus
From Wikipedia, the free encyclopedia
Jump to: navigation, search
Not to be confused with Eudoxus of Cyzicus.

Eudoxus of Cnidus (410 or 408 BC 355 or 347 BC) was a Greek astronomer, mathematician,
scholar and student of Plato. Since all his own works are lost, knowledge of him is obtained from
secondary sources, such as Aratus's poem on astronomy. Theodosius of Bithynia's important
work, Sphaerics, may be based on a work of Eudoxus.

Contents
1 Life

2 Mathematics

3 Astronomy

o 3.1 Eudoxan planetary models


o 3.2 Importance of Eudoxan system

4 Ethics

5 See also

6 References

7 Notes

8 Further reading

9 External links

Life
His name Eudoxus means "honored" or "of good repute" (in Greek , from eu "good" and
doxa "opinion, belief, fame"). It is analogous to the Latin name Benedictus.

Eudoxus's father Aeschines of Cnidus loved to watch stars at night. Eudoxus first travelled to
Tarentum to study with Archytas, from whom he learned mathematics. While in Italy, Eudoxus
visited Sicily, where he studied medicine with Philiston.

Around 387 BC, at the age of 23, he traveled with the physician Theomedon, who according to
Diogenes Lartius some believed was his lover,[1] to Athens to study with the followers of
Socrates. He eventually became the pupil of Plato, with whom he studied for several months, but
due to a disagreement they had a falling out. Eudoxus was quite poor and could only afford an
apartment at the Piraeus. To attend Plato's lectures, he walked the seven miles (11 km) each
direction, each day. Due to his poverty, his friends raised funds sufficient to send him to
Heliopolis, Egypt to pursue his study of astronomy and mathematics. He lived there for 16
months. From Egypt, he then traveled north to Cyzicus, located on the south shore of the Sea of
Marmara, the Propontis. He traveled south to the court of Mausolus. During his travels he
gathered many students of his own.

Around 368 BC, Eudoxus returned to Athens with his students. According to some sources,
around 367 he assumed headship of the Academy during Plato's period in Syracuse, and taught
Aristotle[citation needed]. He eventually returned to his native Cnidus, where he served in the city
assembly. While in Cnidus, he built an observatory and continued writing and lecturing on
theology, astronomy and meteorology. He had one son, Aristagoras, and three daughters, Actis,
Philtis and Delphis.

In mathematical astronomy, his fame is due to the introduction of the astronomical globe, and his
early contributions to understanding the movement of the planets.
His work on proportions shows tremendous insight into numbers; it allows rigorous treatment of
continuous quantities and not just whole numbers or even rational numbers. When it was revived
by Tartaglia and others in the 16th century, it became the basis for quantitative work in science
for a century, until it was replaced by the algebraic methods of Descartes.

Craters on Mars and the Moon are named in his honor. An algebraic curve (the Kampyle of
Eudoxus) is also named after him

a2x4 = b4(x2 + y2).

Mathematics
Eudoxus is considered by some to be the greatest of classical Greek mathematicians, and in all
antiquity, second only to Archimedes. He rigorously developed Antiphon's method of
exhaustion, a precursor to the integral calculus which was also used in a masterly way by
Archimedes in the following century. In applying the method, Eudoxus proved such
mathematical statements as: areas of circles are to one another as the squares of their radii,
volumes of spheres are to one another as the cubes of their radii, the volume of a pyramid is one-
third the volume of a prism with the same base and altitude, and the volume of a cone is one-
third that of the corresponding cylinder.[2]

Eudoxus introduced the idea of non-quantified mathematical magnitude to describe and work
with continuous geometrical entities such as lines, angles, areas and volumes, thereby avoiding
the use of irrational numbers. In doing so, he reversed a Pythagorean emphasis on number and
arithmetic, focusing instead on geometrical concepts as the basis of rigorous mathematics. Some
Pythagoreans, such as Eudoxus' teacher Archytas, had believed that only arithmetic could
provide a basis for proofs. Induced by the need to understand and operate with incommensurable
quantities, Eudoxus established what may have been the first deductive organization of
mathematics on the basis of explicit axioms. The change in focus by Eudoxus stimulated a divide
in mathematics which lasted two thousand years. In combination with a Greek intellectual
attitude unconcerned with practical problems, there followed a significant retreat from the
development of techniques in arithmetic and algebra.[3]

The Pythagoreans had discovered that the diagonal of a square does not have a common unit of
measurement with the sides of the square; this is the famous discovery that the square root of 2
cannot be expressed as the ratio of two integers. This discovery had heralded the existence of
incommensurable quantities beyond the integers and rational fractions, but at the same time it
threw into question the idea of measurement and calculations in geometry as a whole. For
example, Euclid provides an elaborate proof of the Pythagorean theorem (Elements I.47), by
using addition of areas and only much later (Elements VI.31) a simpler proof from similar
triangles, which relies on ratios of line segments.

Ancient Greek mathematicians calculated not with quantities and equations as we do today, but
instead they used proportionalities to express the relationship between quantities. Thus the ratio
of two similar quantities was not just a numerical value, as we think of it today; the ratio of two
similar quantities was a primitive relationship between them.
Eudoxus was able to restore confidence in the use of proportionalities by providing an
astounding definition for the meaning of the equality between two ratios. This definition of
proportion forms the subject of Euclid's Book V.

In Definition 5 of Euclid's Book V we read:

Magnitudes are said to be in the same ratio, the first to the second and the third to the fourth
when, if any equimultiples whatever be taken of the first and third, and any equimultiples
whatever of the second and fourth, the former equimultiples alike exceed, are alike equal to, or
alike fall short of, the latter equimultiples respectively taken in corresponding order.

Let us clarify it by using modern-day notation. If we take four quantities: a, b, c, and d, then the
first and second have a ratio ; similarly the third and fourth have a ratio .

Now to say that we do the following: For any two arbitrary integers, m and n, form
the equimultiples ma and mc of the first and third; likewise form the equimultiples nb and nd
of the second and fourth.

If it happens that ma > nb, then we must also have mc > nd. If it happens that ma = nb, then
we must also have mc = nd. Finally, if it happens that ma < nb, then we must also have mc <
nd.

Notice that the definition depends on comparing the similar quantities ma and nb, and the
similar quantities mc and nd, and does not depend on the existence of a common unit of
measuring these quantities.

The complexity of the definition reflects the deep conceptual and methodological innovation
involved. It brings to mind the famous fifth postulate of Euclid concerning parallels, which is
more extensive and complicated in its wording than the other postulates.

The Eudoxian definition of proportionality uses the quantifier, "for every ..." to harness the
infinite and the infinitesimal, just as do the modern epsilon-delta definitions of limit and
continuity.

Additionally, the Archimedean property stated as definition 4 of Euclid's book V is originally due
not to Archimedes but to Eudoxus.[4]

Astronomy
In ancient Greece, astronomy was a branch of mathematics; astronomers sought to create
geometrical models that could imitate the appearances of celestial motions. Identifying the
astronomical work of Eudoxus as a separate category is therefore a modern convenience. Some
of Eudoxus' astronomical texts whose names have survived include:

Disappearances of the Sun, possibly on eclipses


Oktaeteris (), on an eight-year lunisolar cycle of the calendar

Phaenomena () and Entropon (), on spherical astronomy, probably


based on observations made by Eudoxus in Egypt and Cnidus

On Speeds, on planetary motions

We are fairly well informed about the contents of Phaenomena, for Eudoxus' prose text was the
basis for a poem of the same name by Aratus. Hipparchus quoted from the text of Eudoxus in his
commentary on Aratus.

Eudoxan planetary models

A general idea of the content of On Speeds can be gleaned from Aristotle's Metaphysics XII, 8,
and a commentary by Simplicius of Cilicia (6th century CE) on De caelo, another work by
Aristotle. According to a story reported by Simplicius, Plato posed a question for Greek
astronomers: "By the assumption of what uniform and orderly motions can the apparent motions
of the planets be accounted for?" (quoted in Lloyd 1970, p. 84). Plato proposed that the
seemingly chaotic wandering motions of the planets could be explained by combinations of
uniform circular motions centered on a spherical Earth, apparently a novel idea in the 4th
century.

In most modern reconstructions of the Eudoxan model, the Moon is assigned three spheres:

The outermost rotates westward once in 24 hours, explaining rising and setting.

The second rotates eastward once in a month, explaining the monthly motion of the
Moon through the zodiac.

The third also completes its revolution in a month, but its axis is tilted at a slightly
different angle, explaining motion in latitude (deviation from the ecliptic), and the motion
of the lunar nodes.

The Sun is also assigned three spheres. The second completes its motion in a year instead of a
month. The inclusion of a third sphere implies that Eudoxus mistakenly believed that the Sun had
motion in latitude.

The five visible planets (Venus, Mercury, Mars, Jupiter, and Saturn) are assigned four spheres
each:

The outermost explains the daily motion.

The second explains the planet's motion through the zodiac.


The third and fourth together explain retrogradation, when a planet appears to slow down,
then briefly reverse its motion through the zodiac. By inclining the axes of the two
spheres with respect to each other, and rotating them in opposite directions but with equal
periods, Eudoxus could make a point on the inner sphere trace out a figure-eight shape, or
hippopede.

Importance of Eudoxan system

Callippus, a Greek astronomer of the 4th century, added seven spheres to Eudoxus' original 27
(in addition to the planetary spheres, Eudoxus included a sphere for the fixed stars). Aristotle
described both systems, but insisted on adding "unrolling" spheres between each set of spheres to
cancel the motions of the outer set. Aristotle was concerned about the physical nature of the
system; without unrollers, the outer motions would be transferred to the inner planets.

A major flaw in the Eudoxan system is its inability to explain changes in the brightness of
planets as seen from Earth. Because the spheres are concentric, planets will always remain at the
same distance from Earth. This problem was pointed out in Antiquity by Autolycus of Pitane.
Astronomers responded by introducing the deferent and epicycle, which caused a planet to vary
its distance. However, Eudoxus' importance to Greek astronomy is considerable, as he was the
first to attempt a mathematical explanation of the planets.

Ethics
Aristotle, in The Nicomachean Ethics[5] attributes to Eudoxus an argument in favor of hedonism,
that is, that pleasure is the ultimate good that activity strives for. According to Aristotle, Eudoxus
put forward the following arguments for this position:

1. All things, rational and irrational, aim at pleasure; things aim at what they believe to be
good; a good indication of what the chief good is would be the thing that most things aim
at.

2. Similarly, pleasure's opposite pain is universally avoided, which provides additional


support for the idea that pleasure is universally considered good.

3. People don't seek pleasure as a means to something else, but as an end in its own right.

4. Any other good that you can think of would be better if pleasure were added to it, and it
is only by good that good can be increased.

5. Of all of the things that are good, happiness is peculiar for not being praised, which may
show that it is the crowning good.[6]

Eratosthenes
From Wikipedia, the free encyclopedia
Jump to: navigation, search
This article is about the Greek scholar of the third century BC. For the ancient Athenian
statesman of the fifth century BC, see Eratosthenes (statesman).
Eratosthenes
()

Portrait of Eratosthenes
276 BC
Born
Cyrene
194 BC
Died
Alexandria
Ethnicity Greek
Occupation Scholar, librarian, poet and inventor

Eratosthenes of Cyrene (Ancient Greek: , IPA: [eratostns]; English /r


tsniz/; c. 276 BC[1] c. 195 BC[2]) was a Greek mathematician, geographer, poet,
athlete[citation needed], astronomer, and music theorist.

He was the first person to use the word "geography" in Greek and he invented the discipline of
geography as we understand it.[3] He invented a system of latitude and longitude.

He was the first person to calculate the circumference of the earth by using a measuring system
using stades, or the length of stadiums during that time period (with remarkable accuracy). He
was the first to calculate the tilt of the Earth's axis (also with remarkable accuracy). He may also
have accurately calculated the distance from the earth to the sun and invented the leap day.[4] He
also created the first map of the world incorporating parallels and meridians within his
cartographic depictions based on the available geographical knowledge of the era. In addition,
Eratosthenes was the founder of scientific chronology; he endeavored to fix the dates of the chief
literary and political events from the conquest of Troy.

According to an entry[5] in the Suda (a 10th century reference), his contemporaries nicknamed
him beta, from the second letter of the Greek alphabet, because he supposedly proved himself to
be the second best in the world in almost every field.[6]
Contents
1 Life

2 Eratosthenes' measurement of the Earth's circumference

3 Other astronomical distances

4 Prime numbers

5 Works

6 Things named after Eratosthenes

7 See also

8 Notes

9 Further reading

10 External links

Life

19th century reconstruction of Eratosthenes' map of the known world, c. 194 BC

Eratosthenes was born in Cyrene (in modern-day Libya). He was the third chief librarian of the
Great Library of Alexandria, the center of science and learning in the ancient world, and died in
Alexandria, then the capital of Ptolemaic Egypt.

Eratosthenes studied in Alexandria, and claimed to have also studied for some years in Athens.
In 236 BC, he was appointed by Ptolemy III Euergetes as librarian of the Alexandrian library,
succeeding the second librarian, Apollonius of Rhodes.[7] He made several important
contributions to mathematics and science, and was a good friend to Archimedes. Around
255 BC, he invented the armillary sphere. In On the Circular Motions of the Celestial Bodies,
Cleomedes credited him with having calculated the Earth's circumference around 240 BC, using
knowledge of the angle of elevation of the sun at noon on the summer solstice in Alexandria and
on Elephantine Island near Syene (now Aswan, Egypt).

Eratosthenes criticized Aristotle for arguing that humanity was divided into Greeks and
barbarians, and that the Greeks should keep themselves racially pure, believing there was good
and bad in every nation.[8] By 195 BC, Eratosthenes became blind. He died in 194 BC at the age
of 82.

Eratosthenes' measurement of the Earth's circumference

Illustration showing a portion of the globe showing a part of the African continent. The sun
beams shown as 2 rays hitting earth at Syene and Alexandria. Angle of sun beam and the
gnomons (vertical sticks) is shown at Alexandria which allowed Eratosthenes' estimates of radius
and circumference of Earth.

Measurements taken at Alexandria (A) and Syene (S)

Eratosthenes calculated the circumference of the Earth without leaving Egypt. Eratosthenes knew
that, on the summer solstice, at local noon in the Ancient Egyptian city of Swenet (known in
Greek as Syene, and in the modern day as Aswan) on the Tropic of Cancer, the sun would appear
at the zenith, directly overhead (he had been told that the shadow of someone looking down a
deep well would block the reflection of the Sun at noon). Using a gnomon, he measured the sun's
angle of elevation at noon on the solstice in his hometown of Alexandria, and found it to be
1/50th of a circle (712') south of the zenith. Assuming that the Earth was spherical (360), and
that Alexandria was due north of Syene, he concluded that the meridian arc distance from
Alexandria to Syene must therefore be 1/50 = 712'/360, and was therefore 1/50 of the total
circumference of the Earth. His knowledge of the size of Egypt after many generations of
surveying trips for the Pharaonic bookkeepers gave a distance between the cities of 5,000 stadia
(about 500 geographical miles or 927.7 km). This distance was corroborated by inquiring about
the time that it takes to travel from Syene to Alexandria by camel. He rounded the result to a
final value of 700 stadia per degree, which implies a circumference of 252,000 stadia. The exact
size of the stadion he used is frequently argued. The common Attic stadion was about 185 m,[9]
which would imply a circumference of 46,620 km, which is off the actual circumference by
16.3%; too large an error to be considered as 'accurate'. However, if we assume that Eratosthenes
used the "Egyptian stadion"[10] of about 157.5 m, his measurement turns out to be 39,690 km, an
error of less than 2%.[11]

Other astronomical distances

This section needs additional citations for verification. (March 2009)

Eusebius of Caesarea in his Preparatio Evangelica includes a brief chapter of three sentences on
celestial distances (Book XV, Chapter 53). He states simply that Eratosthenes found the distance
to the sun to be " " (literally "of stadia myriads
400 and 80,000") and the distance to the moon to be 780,000 stadia. The expression for the
distance to the sun has been translated either as 4,080,000 stadia (1903 translation by E. H.
Gifford), or as 804,000,000 stadia (edition of Edouard des Places, dated 19741991). The
meaning depends on whether Eusebius meant 400 myriad plus 80,000 or "400 and 80,000"
myriad. With a stadium of 185 meters, 804,000,000 stadia is 149,000,000 kilometers,
approximately the distance from the earth to the sun.

Prime numbers
Main articles: Sieve of Eratosthenes and primality test

Eratosthenes also proposed a simple algorithm for finding prime numbers. This algorithm is
known in mathematics as the Sieve of Eratosthenes.

In mathematics, the sieve of Eratosthenes (Greek: ), one of a number of


prime number sieves, is a simple, ancient algorithm for finding all prime numbers up to any
given limit. It does so by iteratively marking as composite (i.e. not prime) the multiples of each
prime, starting with the multiples of 2. The multiples of a given prime are generated starting
from that prime, as a sequence of numbers with the same difference, equal to that prime, between
consecutive numbers. This is the sieve's key distinction from using trial division to sequentially
test each candidate number for divisibility by each prime.

Works
(On the Measurement of the Earth)[12] (lost, summarized
by Cleomedes)

Geographica (lost, criticized by Strabo)

Arsinoe (a memoir of queen Arsinoe; lost; quoted by Athenaeus in the Deipnosophistae)

A fragmentary collection of Hellenistic myths about the constellations, called


Catasterismi (Katasterismoi), was attributed to Eratosthenes, perhaps to add to its
credibility.

Things named after Eratosthenes

Eratosthenes (crater) on the moon

Eratosthenes (crater) on the moon

Eratosthenian period in the lunar geologic timescale

Eratosthenes Seamount in the eastern Mediterranean Sea

Who Was Eratosthenes?


Eratosthenes (276 BC-194 BC) was a Greek mathematician, geographer and astronomer. He was
born in Cyrene (now Libya) and died in Ptolemaic Alexandria. He is noted for devising a map
system based on latitude and longitude lines and computing the size of the Earth.

Eratosthenes studied at Alexandria and for some years in Athens. In 236 BC he was appointed by
Ptolemy III Euergetes I as librarian of the Alexandrian library. He made several important
contributions to mathematics and science, and was a good friend to Archimedes. Around 255 BC
he invented the armillary sphere (an astronomical instrument for determining celestial positions),
which was widely used until the invention of the orrery in the 18th century.

Circa 200 BC Eratosthenes is thought to have coined or to have adopted the word geography, the
descriptive study of the Earth.

Eratosthenes' other contributions include:

The Sieve of Eratosthenes as a way of finding prime numbers.

The measurement of the Sun-Earth distance, now called the astronomical unit
(804,000,000 stadia, 1 stadion varies from 157 to 209 meter).

The measurement of the distance to the Moon (780,000 stadia).

The measurement of the inclination of the ecliptic with an angle error of 7'.

He compiled a star catalogue containing 675 stars, which was not preserved.

A map of the Nile's route as far as Khartoum.

A map of the entire known world, from the British Isles to Ceylon, and from the Caspian
Sea to Ethiopia.

Eratosthenes' Experiment
Eratosthenes will always be remembered for the calculation of the Earth's circumference circa
240 BC, using trigonometry and knowledge of the angle of elevation of the Sun at noon in
Alexandria and Syene (now Aswan, Egypt). The calculation is based on the assumption that the
Earth is spherical and that the Sun is so far away that its rays can be taken as parallel.

Details of his method he published in a work On the measurement of the Earth which
unfortunately was lost. We know indirectly about his method from other authors.

Before we begin a few definitions:

Tropic of Cancer - is one of five major circles of latitude that mark maps of the Earth. The
Tropic of Cancer currently latitude is 23 26 22 north of the Equator.

Local noon is when the sun is the highest in the sky and can be quite different from 12:00 noon
on the clock.

Solstice - is an astronomical event that happens twice each year, when the tilt of the Earth's axis
is most inclined toward or away from the Sun. In the northern hemisphere, the maximum
inclination toward the sun is around 21 June (the summer solstice) and with the maximum
inclination away around 21 December (the winter solstice). For the southern hemisphere winter
and summer solstices are exchanged.
What matters for our experiment is the fact that on the summer solstice, local noon, the sun rays
are just overhead (at a right angle to the ground) on the Tropic of Cancer.

Eratosthenes' Experiment

Eratosthenes knew that on the summer solstice at local noon on the Tropic of Cancer, the Sun
would appear at the zenith, directly overhead (sun elavation of 90) - though Syene was in fact
slightly north of the tropic. He also knew, from using a vertical stick and measuring the cast
shadow, that in his hometown of Alexandria, the angle of elevation of the Sun would be 83 or 7
south of the zenith at the same time. Assuming that Alexandria was due north of Syene -
Alexandria is in fact on a more westerly longitude - he concluded, using geometry of parallel
lines, that the distance from Alexandria to Syene must be 7/360 of the total circumference of the
Earth. The distance between the cities was known from caravan travellings to be about 5,000
stadia. He established a final value of 700 stadia per degree, which implies a circumference of
252,000 stadia. The exact size of the stadion he used is no longer known (the common Attic
stadion was about 185 m), but it is generally believed that Eratosthenes' value corresponds to
between 39,690 km and 46,620 km. The circumference of the Earth around the poles is now
measured at around 40,008 km. Eratosthenes result is not bad at all.

Very interesting is that the measurement of the distance between Alexandria and Syene is based
on the estimated average speed of a caravan of camels that traveled this distance(!). Camels
traveled the distance many times to obtain an average estimate of the distance. Whether this is
true is not clear.

Repeat Eratosthenes' Experiment


Eratosthenes measured, at his local noon in Alexandria, the angle of elevation of the sun on the
summer solstice (21 June). Eratosthenes used the local noon and no other time of the day since at
local noon all relevant places and sunrays are placed on the same imaginary plane enabling the
use of simple geometry for his calculations. In order to repeat Eratosthenes experiment youll
have to do the same.

First, calculate your local noon because it may be quite different from 12:00 noon on the clock.
There are several ways to compute its exact occurrence. Basically, local noon is half-way
between sunrise and sunset. You can obtain sunrise and sunset times, for June 21, from your local
paper or from this link: http://aa.usno.navy.mil/data/docs/RS_OneDay.php which also calculates
local noon (sun transit). You can also obtain it by yourself by using a sundial or find out when
the shadow is the shortest around noon time.

On June 21 erect a vertical straight stick or pole of about 1 meter using a carpenters level and
measure the length of the shadow it casts at your local noon. With simple trigonometry you can
obtain the angle of the elevation of the sun. You can also obtain the angle, without trigonometry,
by drawing the stick and shadow proportionally and measuring it with a protractor. You can
compare your results with a web based applet like this: http://www.jgiesen.de/azimuth but be
careful to use it correctly (insert your correct time zone, local noon, coordinates, date and ensure
that the dropdown menu points to elevation).

After you get the angle of sun elevation, its very easy to calculate the zenith angle by subtracting
it from 90, like Eratosthenes did. Now youll have to measure the distance from your location to
the Tropic of Cancer latitude line - not by camel caravans of course, the Eratosthenes way. You
can use a relatively large scale map, but take in account that maps tend to distort distance and the
best option is to use a globe. The distance from your location to the Tropic of Cancer should be
measured from north to south. In other words the distance line has to cut the Tropic of Cancer at
a right angle. There are also web based calculators for this:
http://facstaff.gpc.edu/~pgore/ISCI/earthcircumference.html.
Now it's easy to calculate the Earth circumference by using the following formula:

Likewise, you can also perform this experiment on the winter solstice that takes place around 21
December, but youll have to measure your distance from the Tropic of Capricorn instead from
the Tropic of Cancer because on this date the sun reaches its highest degree of elevation on the
Tropic of Capricorn (23 26 22 south of the Equator).

It is also possible to perform this experiment on the two Equinoxes which occur on 20 March
and 23 September each year when the sun is crossing the equator at the local noon on those dates
and the sun rays are just overhead the equator at a right angle to the ground. But instead to
measure your distance from the Tropic of Cancer or the Tropic of Capricorn youll have to
measure it from the equator.

There is another option and you can perform this experiment on any other date of the year, at
local noon time, but you should have some partner located on your longitude willing to measure
sun elevation at the same time. Take in account that you'll have to be a little careful treating
correctly the sun angles obtained in this case.

At any date the sun reaches its highest position, at noon time, at some latitude. From here is clear
that if the two places involved are located on the same side of this latitude (north or south) the
shadows will be casted at the same direction and the obtained angles should be subtracted from
each other, whereas if the places are located on different sides of this latitude the shadows will be
casted at different directions (southward or northward) and the angles should be added up.

Further Reading

Diophantus
From Wikipedia, the free encyclopedia
Jump to: navigation, search
For the general, see Diophantus (general).
Title page of the 1621 edition of Diophantus' Arithmetica, translated into Latin by Claude
Gaspard Bachet de Mziriac.

Diophantus of Alexandria (Ancient Greek: . b. between A.D. 200


and 214, d. between 284 and 298 at age 84), sometimes called "the father of algebra", was an
Alexandrian Greek mathematician[1][2][3][4] and the author of a series of books called Arithmetica.
These texts deal with solving algebraic equations, many of which are now lost. In studying
Arithmetica, Pierre de Fermat concluded that a certain equation considered by Diophantus had no
solutions, and noted without elaboration that he had found "a truly marvelous proof of this
proposition," now referred to as Fermat's Last Theorem. This led to tremendous advances in
number theory, and the study of Diophantine equations ("Diophantine geometry") and of
Diophantine approximations remain important areas of mathematical research. Diophantus
coined the term to refer to an approximate equality.[5] This term was rendered as
adaequalitat in Latin, and became the technique of adequality developed by Pierre de Fermat to
find maxima for functions and tangent lines to curves. Diophantus was the first Greek
mathematician who recognized fractions as numbers; thus he allowed positive rational numbers
for the coefficients and solutions. In modern use, Diophantine equations are usually algebraic
equations with integer coefficients, for which integer solutions are sought. Diophantus also made
advances in mathematical notation.

Contents
1 Biography

2 Arithmetica
o 2.1 History

o 2.2 Margin writing by Fermat and Chortasmenos

3 Other works

o 3.1 The Porisms

o 3.2 Polygonal numbers and geometric elements

4 Influence

o 4.1 The father of algebra?

5 Diophantine analysis

6 Mathematical notation

7 See also

8 Notes

9 References

10 Further reading

11 External links

Biography
Little is known about the life of Diophantus. He lived in Alexandria, Egypt, probably from
between A.D. 200 and 214 to 284 or 298. Much of our knowledge of the life of Diophantus is
derived from a 5th century Greek anthology of number games and strategy puzzles. One of the
problems (sometimes called his epitaph) states:

'Here lies Diophantus,' the wonder behold.


Through art algebraic, the stone tells how old:
'God gave him his boyhood one-sixth of his life,
One twelfth more as youth while whiskers grew rife;
And then yet one-seventh ere marriage begun;
In five years there came a bouncing new son.
Alas, the dear child of master and sage
After attaining half the measure of his father's life chill fate took him. After consoling his
fate by the science of numbers for four years, he ended his life.'

This puzzle implies that Diophantus lived to be 84 years old. However, the accuracy of the
information cannot be independently confirmed.

In popular culture, this puzzle was the Puzzle No.142 in Professor Layton and Pandora's Box as
one of the hardest solving puzzles in the game, which needed to be unlocked by solving other
puzzles first.

Arithmetica
See also: Arithmetica

The Arithmetica is the major work of Diophantus and the most prominent work on algebra in
Greek mathematics. It is a collection of problems giving numerical solutions of both determinate
and indeterminate equations. Of the original thirteen books of which Arithmetica consisted only
six have survived, though there are some who believe that four Arab books discovered in 1968
are also by Diophantus.[6] Some Diophantine problems from Arithmetica have been found in
Arabic sources.

It should be mentioned here that Diophantus never used general methods in his solutions.
Hermann Hankel, renowned German mathematician made the following remark regarding
Diophantus.

Our author (Diophantos) not the slightest trace of a general, comprehensive method is
discernible; each problem calls for some special method which refuses to work even for the most
closely related problems. For this reason it is difficult for the modern scholar to solve the 101st
problem even after having studied 100 of Diophantoss solutions [7][dubious discuss]

History

Like many other Greek mathematical treatises, Diophantus was forgotten in Western Europe
during the so-called Dark Ages, since the study of ancient Greek had greatly declined. The
portion of the Greek Arithmetica that survived, however, was, like all ancient Greek texts
transmitted to the early modern world, copied by, and thus known to, medieval Byzantine
scholars. In addition, some portion of the Arithmetica probably survived in the Arab tradition
(see above). In 1463 German mathematician Regiomontanus wrote:

No one has yet translated from the Greek into Latin the thirteen books of Diophantus, in
which the very flower of the whole of arithmetic lies hidden . . . .

Arithmetica was first translated from Greek into Latin by Bombelli in 1570, but the translation
was never published. However, Bombelli borrowed many of the problems for his own book
Algebra. The editio princeps of Arithmetica was published in 1575 by Xylander. The best known
Latin translation of Arithmetica was made by Bachet in 1621 and became the first Latin edition
that was widely available. Pierre de Fermat owned a copy, studied it, and made notes in the
margins.

Margin writing by Fermat and Chortasmenos

Problem II.8 in the Arithmetica (edition of 1670), annotated with Fermat's comment which
became Fermat's Last Theorem.

The 1621 edition of Arithmetica by Bachet gained fame after Pierre de Fermat wrote his famous
"Last Theorem" in the margins of his copy:

If an integer n is greater than 2, then has no solutions in non-zero


integers a, b, and c. I have a truly marvelous proof of this proposition which this margin
is too narrow to contain.

Fermat's proof was never found, and the problem of finding a proof for the theorem went
unsolved for centuries. A proof was finally found in 1994 by Andrew Wiles after working on it
for seven years. It is believed that Fermat did not actually have the proof he claimed to have.
Although the original copy in which Fermat wrote this is lost today, Fermat's son edited the next
edition of Diophantus, published in 1670. Even though the text is otherwise inferior to the 1621
edition, Fermat's annotationsincluding the "Last Theorem"were printed in this version.

Fermat was not the first mathematician so moved to write in his own marginal notes to
Diophantus; the Byzantine scholar John Chortasmenos (14th/15th C.) had written "Thy soul,
Diophantus, be with Satan because of the difficulty of your theorems" next to the same problem.

Other works
Diophantus wrote several other books besides Arithmetica, but very few of them have survived.

The Porisms

Diophantus himself refers to a work which consists of a collection of lemmas called The Porisms
(or Porismata), but this book is entirely lost. Some scholars think that The porisms may have
actually been a section of Arithmetica that is now lost.[citation needed]

Although The Porisms is lost, we know three lemmas contained there, since Diophantus refers to
them in the Arithmetica. One lemma states that the difference of the cubes of two rational
numbers is equal to the sum of the cubes of two other rational numbers, i.e. given any a and b,
with a > b, there exist c and d, all positive and rational, such that

Polygonal numbers and geometric elements

Diophantus is also known to have written on polygonal numbers, a topic of great interest to
Pythagoras and Pythagoreans. Fragments of a book dealing with polygonal numbers are extant.

A book called Preliminaries to the Geometric Elements has been traditionally attributed to Hero
of Alexandria. It has been studied recently by Wilbur Knorr, who suggested that the attribution to
Hero is incorrect, and that the true author is Diophantus.[8]

Influence
Diophantus' work has had a large influence in history. Editions of Arithmetica exerted a profound
influence on the development of algebra in Europe in the late sixteenth and through the
seventeenth and eighteenth centuries. Diophantus and his works have also influenced Arab
mathematics and were of great fame among Arab mathematicians. Diophantus' work created a
foundation for work on algebra and in fact much of advanced mathematics is based on algebra.
As far as we know Diophantus did not affect the lands of the Orient much and how much he
affected India is a matter of debate.

The father of algebra?

Diophantus is often called the father of algebra" because he contributed greatly to number
theory, mathematical notation, and because Arithmetica contains the earliest known use of
syncopated notation.[9] However, it seems that many of the methods for solving linear and
quadratic equations used by Diophantus go back to Babylonian mathematics. For this, and other,
reasons mathematical historian Kurt Vogel writes: Diophantus was not, as he has often been
called, the father of algebra. Nevertheless, his remarkable, if unsystematic, collection of
indeterminate problems is a singular achievement that was not fully appreciated and further
developed until much later.[10]
Diophantine analysis
See also: Diophantine equation

Today Diophantine analysis is the area of study where integer (whole number) solutions are
sought for equations, and Diophantine equations are polynomial equations with integer
coefficients to which only integer solutions are sought. It is usually rather difficult to tell whether
a given Diophantine equation is solvable. Most of the problems in Arithmetica lead to quadratic
equations. Diophantus looked at 3 different types of quadratic equations: ,
, and . The reason why there were three cases to Diophantus,
while today we have only one case, is that he did not have any notion for zero and he avoided
negative coefficients by considering the given numbers to all be positive in each of the
three cases above. Diophantus was always satisfied with a rational solution and did not require a
whole number which means he accepted fractions as solutions to his problems. Diophantus
considered negative or irrational square root solutions "useless", "meaningless", and even
"absurd". To give one specific example, he calls the equation 'absurd' because it
would lead to a negative value for x. One solution was all he looked for in a quadratic equation.
There is no evidence that suggests Diophantus even realized that there could be two solutions to
a quadratic equation. He also considered simultaneous quadratic equations.

Mathematical notation
Diophantus made important advances in mathematical notation. He was the first person to use
algebraic notation and symbolism. Before him everyone wrote out equations completely.
Diophantus introduced an algebraic symbolism that used an abridged notation for frequently
occurring operations, and an abbreviation for the unknown and for the powers of the unknown.
Mathematical historian Kurt Vogel states:

The symbolism that Diophantus introduced for the first time, and undoubtedly devised himself,
provided a short and readily comprehensible means of expressing an equation... Since an
abbreviation is also employed for the word equals, Diophantus took a fundamental step from
verbal algebra towards symbolic algebra.[citation needed]

Although Diophantus made important advances in symbolism, he still lacked the necessary
notation to express more general methods. This caused his work to be more concerned with
particular problems rather than general situations. Some of the limitations of Diophantus'
notation are that he only had notation for one unknown and, when problems involved more than
a single unknown, Diophantus was reduced to expressing "first unknown", "second unknown",
etc. in words. He also lacked a symbol for a general number n. Where we would write
, Diophantus has to resort to constructions like : ... a sixfold number
increased by twelve, which is divided by the difference by which the square of the number
exceeds three.
Algebra still had a long way to go before very general problems could be written down and
solved succinctly.

See also
ErdsDiophantine graph

Diophantus II.VIII

Polynomial Diophantine equation

Notes
1. ^ Research Machines plc. (2004). The Hutchinson dictionary of scientific
biography. Abingdon, Oxon: Helicon Publishing. p. 312. "Diophantus (lived c.A.D. 270-
280) Greek mathematician who, in solving linear mathematical problems, developed an
early form of algebra."

2. ^ Boyer, Carl B. (1991). "Revival and Decline of Greek Mathematics". A History


of Mathematics (Second ed.). John Wiley & Sons, Inc.. p. 178. ISBN 0-471-54397-7. "At
the beginning of this period, also known as the Later Alexandrian Age, we find the
leading Greek algebraist, Diophantus of Alexandria, and toward its close there appeared
the last significant Greek geometer, Pappus of Alexandria."

3. ^ Cooke, Roger (1997). "The Nature of Mathematics". The History of


Mathematics: A Brief Course. Wiley-Interscience. p. 7. ISBN 0-471-18082-3. "Some
enlargement in the sphere in which symbols were used occurred in the writings of the
third-century Greek mathematician Diophantus of Alexandria, but the same defect was
present as in the case of Akkadians."

4. ^ Victor J. Katz (1998). A History of Mathematics: An Introduction, p. 184.


Addison Wesley, ISBN 0-321-01618-1.

"But what we really want to know is to what extent the Alexandrian mathematicians of
the period from the first to the fifth centuries C.E. were Greek. Certainly, all of them
wrote in Greek and were part of the Greek intellectual community of Alexandria. And
most modern studies conclude that the Greek community coexisted [...] So should we
assume that Ptolemy and Diophantus, Pappus and Hypatia were ethnically Greek, that
their ancestors had come from Greece at some point in the past but had remained
effectively isolated from the Egyptians? It is, of course, impossible to answer this
question definitively. But research in papyri dating from the early centuries of the
common era demonstrates that a significant amount of intermarriage took place between
the Greek and Egyptian communities [...] And it is known that Greek marriage contracts
increasingly came to resemble Egyptian ones. In addition, even from the founding of
Alexandria, small numbers of Egyptians were admitted to the privaleged classes in the
city to fulfill numerous civic roles. Of course, it was essential in such cases for the
Egyptians to become "Hellenized," to adopt Greek habits and the Greek language. Given
that the Alexandrian mathematicians mentioned here were active several hundred years
after the founding of the city, it would seem at least equally possible that they were
ethnically Egyptian as that they remained ethnically Greek. In any case, it is unreasonable
to portray them with purely European features when no physical descriptions exist."

5. ^ Katz, Mikhail G.; Schaps, David; Shnider, Steve (2013), "Almost Equal: The
Method of Adequality from Diophantus to Fermat and Beyond", Perspectives on Science
21 (3), arXiv:1210.7750

6. ^ J. Sesiano (1982). Books IV to VII of Diophantus' Arithmetica in the Arabic


Translation Attributed to Qusta ibn Luqa. New York/Heidelberg/Berlin: Springer-Verlag.
p. 502.

7. ^ Hankel H., Geschichte der mathematic im altertum und mittelalter, Leipzig,


1874. (translated to English by Ulrich Lirecht in Chinese Mathematics in the thirteenth
century, Dover publications, New York, 1973.

8. ^ Knorr, Wilbur: Arithmtike stoicheisis: On Diophantus and Hero of


Alexandria, in: Historia Matematica, New York, 1993, Vol.20, No.2, 180-192

9. ^ Carl B. Boyer, A History of Mathematics, Second Edition (Wiley, 1991), page


228

^ Harald Kittel, bersetzung: ein internationales Handbuch zur


bersetzungsforschung, VolDiophantus II.VIII

From Wikipedia, the free encyclopedia


Jump to: navigation, search
Diophantus II.VIII: Intersection of the line CB and the circle gives a rational point (x0,y0).

The eighth problem of the second book of Diophantus's Arithmetica is to divide a square into
a sum of two squares.

Contents
1 The solution given by Diophantus

2 Geometrical interpretation

3 Generalization of Diophantus's solution

4 See also

5 References

The solution given by Diophantus


Diophantus takes the square to be 16 and solves the problem as follows:[1]

To divide a given square into a sum of two squares.

To divide 16 into a sum of two squares.

Let the first summand be , and thus the second . The latter is to be a square. I form
the square of the difference of an arbitrary multiple of x diminished by the root [of] 16, that is,
diminished by 4. I form, for example, the square of 2x 4. It is . I put this
expression equal to . I add to both sides and subtract 16. In this way I
obtain , hence .

Thus one number is 256/25 and the other 144/25. The sum of these numbers is 16 and each
summand is a square.

Geometrical interpretation
Geometrically, we may illustrate this method by drawing the circle x2 + y2 = 42 and the line
y = 2x - 4. The pair of squares sought are then x02 and y02, where (x0, y0) is the point not on the y-
axis where the line and circle intersect. This is shown in the diagram to the right.
Generalization of Diophantus's solution

Diophantus II.VIII: Generalized solution in which the sides of triangle OAB form a rational
triple if line CB has a rational gradient t.

We may generalize Diophantus's solution to solve the problem for any given square, which we
will represent algebraically as a2. Also, since Diophantus refers to an arbitrary multiple of x, we
will take the arbitrary multiple to be tx. Then:

Therefore, we find that one of the summands is and the other is . The sum
of these numbers is and each summand is a square. Geometrically, we have intersected the
circle x2 + y2 = a2 with the line y = tx - a, as shown in the diagram to the right.[2] Writing the
lengths, OB, OA, and AB, of the sides of triangle OAB as an ordered tuple, we obtain the triple

.
The specific result obtained by Diophantus may be obtained by taking a = 4 and t = 2:

We see that Diophantus' particular solution is in fact a subtly disguised (3, 4, 5) triple. However,
as the triple will always be rational as long as a and t are rational, we can obtain an infinity of
rational triples by changing the value of t, and hence changing the value of the arbitrary multiple
of x.

This algebraic solution needs only one additional step to arrive at the Platonic sequence
and that is to multiply all sides of the above triple by a factor . Notice also
that if a = 1, the sides [OB, OA, AB] reduce to

In modern notation this is just written in terms of the cotangent of /2, which
in the particular example given by Diophantus has a value of 2, the multiplier in the arbitrary
multiple of x. Upon clearing denominators, this expression will generate Pythagorean triples.
Intriguingly, the arbitrary multiple of x has become the cornerstone of the generator
expression(s).

Diophantus II.IX reaches the same solution by an even quicker route which is very similar to the
'generalized solution' above. Once again the problem is to divide 16 into two squares.[3]

Let the first number be N and the second an arbitrary multiple of N diminished by the root (of)
16. For example 2N 4. Then:

Historical note: Fermat's famous comment which later became Fermat's Last Theorem appears
sandwiched between 'Quaestio VIII' and 'Quaestio IX' on page 61 of a 1670 edition of Ar

10. ume 2 p. 1123, 1124


Fermat's Last Theorem
From Wikipedia, the free encyclopedia
Jump to: navigation, search
For other theorems named after Pierre de Fermat, see Fermat's theorem.

The 1670 edition of Diophantus' Arithmetica includes Fermat's commentary, particularly his
"Last Theorem" (Observatio Domini Petri de Fermat).
In number theory, Fermat's Last Theorem (sometimes called Fermat's conjecture, especially
in older texts) states that no three positive integers a, b, and c can satisfy the equation an + bn = cn
for any integer value of n greater than two.

This theorem was first conjectured by Pierre de Fermat in 1637, famously in the margin of a
copy of Arithmetica where he claimed he had a proof that was too large to fit in the margin. No
successful proof was published until 1995 despite the efforts of countless mathematicians during
the 358 intervening years. The unsolved problem stimulated the development of algebraic
number theory in the 19th century and the proof of the modularity theorem in the 20th century. It
is among the most famous theorems in the history of mathematics and prior to its 1995 proof was
in the Guinness Book of World Records for "most difficult mathematical problems".

Contents
1 History

2 Mathematical context

o 2.1 Pythagorean triples

o 2.2 Diophantine equations

3 Fermat's conjecture

4 Proofs for specific exponents

5 Sophie Germain

6 Ernst Kummer and the theory of ideals

7 Mordell conjecture

8 Computational studies

9 Connection with elliptic curves

10 Wiles's general proof

11 Exponents other than positive integers

o 11.1 Rational exponents


o 11.2 Negative exponents

11.2.1 n = 1

11.2.2 n = 2

11.2.3 Integer n < 2

12 Did Fermat possess a general proof?

13 Monetary prizes

14 In popular culture

15 See also

16 Notes

17 References

18 Bibliography

19 Further reading

20 External links

History
Fermat left no proof of the conjecture for all n, but he did prove the special case n = 4. This
reduced the problem to proving the theorem for exponents n that are prime numbers. Over the
next two centuries (16371839), the conjecture was proven for only the primes 3, 5, and 7,
although Sophie Germain proved a special case for all primes less than 100. In the mid-19th
century, Ernst Kummer proved the theorem for regular primes. Building on Kummer's work and
using sophisticated computer studies, other mathematicians were able to prove the conjecture for
all odd primes up to four million.

The final proof of the conjecture for all n came in the late 20th century. In 1984, Gerhard Frey
suggested the approach of proving the conjecture through a proof of the modularity theorem for
elliptic curves. Building on work of Ken Ribet, Andrew Wiles succeeded in proving enough of
the modularity theorem to prove Fermat's Last Theorem, with the assistance of Richard Taylor.
Wiles's achievement was reported widely in the popular press, and has been popularized in books
and television programs.
Mathematical context
Pythagorean triples

Main article: Pythagorean triple

Pythagorean triples are a set of three integers (a, b, c) that satisfy a special case of Fermat's
equation (n = 2)[1]

Examples of Pythagorean triples include (3, 4, 5) and (5, 12, 13). There are infinitely many such
triples,[2] and methods for generating such triples have been studied in many cultures, beginning
with the Babylonians[3] and later ancient Greek, Chinese, and Indian mathematicians.[4] The
traditional interest in Pythagorean triples connects with the Pythagorean theorem;[5] in its
converse form, it states that a triangle with sides of lengths a, b, and c has a right angle between
the a and b legs when the numbers are a Pythagorean triple. Right angles have various practical
applications, such as surveying, carpentry, masonry, and construction. Fermat's Last Theorem is
an extension of this problem to higher powers, stating that no solution exists when the exponent
2 is replaced by any larger integer.

Diophantine equations

Main article: Diophantine equation

Fermat's equation xn + yn = zn is an example of a Diophantine equation.[6] A Diophantine equation


is a polynomial equation in which the solutions must be integers.[7] Their name derives from the
3rd-century Alexandrian mathematician, Diophantus, who developed methods for their solution.
A typical Diophantine problem is to find two integers x and y such that their sum, and the sum of
their squares, equal two given numbers A and B, respectively:

Diophantus's major work is the Arithmetica, of which only a portion has survived.[8] Fermat's
conjecture of his Last Theorem was inspired while reading a new edition of the Arithmetica,[9]
which was translated into Latin and published in 1621 by Claude Bachet.[10]

Diophantine equations have been studied for thousands of years. For example, the solutions to
the quadratic Diophantine equation x2 + y2 = z2 are given by the Pythagorean triples, originally
solved by the Babylonians (c. 1800 BC).[11] Solutions to linear Diophantine equations, such as
26x + 65y = 13, may be found using the Euclidean algorithm (c. 5th century BC).[12] Many
Diophantine equations have a form similar to the equation of Fermat's Last Theorem from the
point of view of algebra, in that they have no cross terms mixing two letters, without sharing its
particular properties. For example, it is known that there are infinitely many positive integers x,
y, and z such that xn + yn = zm where n and m are relatively prime natural numbers.[note 1]

Fermat's conjecture

Problem II.8 in the 1621 edition of the Arithmetica of Diophantus. On the right is the famous
margin which was too small to contain Fermat's alleged proof of his "last theorem".

Problem II.8 of the Arithmetica asks how a given square number is split into two other squares;
in other words, for a given rational number k, find rational numbers u and v such that k2 = u2 + v2.
Diophantus shows how to solve this sum-of-squares problem for k = 4 (the solutions being u =
16/5 and v = 12/5).[13]

Around 1637, Fermat wrote his Last Theorem in the margin of his copy of the Arithmetica next
to Diophantus' sum-of-squares problem:[14]

Cubum autem in duos cubos, aut it is impossible to separate a cube into


quadratoquadratum in duos two cubes, or a fourth power into two
quadratoquadratos, et generaliter nullam in fourth powers, or in general, any
infinitum ultra quadratum potestatem in duos power higher than the second, into
eiusdem nominis fas est dividere cuius rei two like powers. I have discovered a
demonstrationem mirabilem sane detexi. Hanc truly marvelous proof of this, which
marginis exiguitas non caperet. this margin is too narrow to contain.
[15]
Although Fermat's general proof is unknown, his proof of one case (n = 4) by infinite descent has
survived.[16] Fermat posed the cases of n = 4 and of n = 3 as challenges to his mathematical
correspondents, such as Marin Mersenne, Blaise Pascal, and John Wallis.[17] However, in the last
thirty years of his life, Fermat never again wrote of his "truly marvellous proof" of the general
case.

After Fermat's death in 1665, his son Clment-Samuel Fermat produced a new edition of the
book (1670) augmented with his father's comments.[18] The margin note became known as
Fermat's Last Theorem,[19] as it was the last of Fermat's asserted theorems to remain unproven.[20]

Proofs for specific exponents


Main article: Proof of Fermat's Last Theorem for specific exponents

Only one mathematical proof by Fermat has survived, in which Fermat uses the technique of
infinite descent to show that the area of a right triangle with integer sides can never equal the
square of an integer.[21] His proof is equivalent to demonstrating that the equation

has no primitive solutions in integers (no pairwise coprime solutions). In turn, this proves
Fermat's Last Theorem for the case n=4, since the equation a4 + b4 = c4 can be written as c4 b4 =
(a2)2.

Alternative proofs of the case n = 4 were developed later[22] by Frnicle de Bessy (1676),[23]
Leonhard Euler (1738),[24] Kausler (1802),[25] Peter Barlow (1811),[26] Adrien-Marie Legendre
(1830),[27] Schopis (1825),[28] Terquem (1846),[29] Joseph Bertrand (1851),[30] Victor Lebesgue
(1853, 1859, 1862),[31] Theophile Pepin (1883),[32] Tafelmacher (1893),[33] David Hilbert (1897),
[34]
Bendz (1901),[35] Gambioli (1901),[36] Leopold Kronecker (1901),[37] Bang (1905),[38] Sommer
(1907),[39] Bottari (1908),[40] Karel Rychlk (1910),[41] Nutzhorn (1912),[42] Robert Carmichael
(1913),[43] Hancock (1931),[44] and Vrnceanu (1966).[45]

For another proof for n=4 by infinite descent, see Infinite descent: Non-solvability of r2 + s4 = t4.
For various proofs for n=4 by infinite descent, see Grant and Perella (1999),[46] Barbara (2007),[47]
and Dolan (2011).[48]

After Fermat proved the special case n = 4, the general proof for all n required only that the
theorem be established for all odd prime exponents.[49] In other words, it was necessary to prove
only that the equation an + bn = cn has no integer solutions (a, b, c) when n is an odd prime
number. This follows because a solution (a, b, c) for a given n is equivalent to a solution for all
the factors of n. For illustration, let n be factored into d and e, n = de. The general equation

an + bn = cn

implies that (ad, bd, cd) is a solution for the exponent e


(ad)e + (bd)e = (cd)e.

Thus, to prove that Fermat's equation has no solutions for n > 2, it suffices to prove that it has no
solutions for at least one prime factor of every n. All integers n > 2 contain a factor of 4, or an
odd prime number, or both. Therefore, Fermat's Last Theorem can be proven for all n if it can be
proven for n = 4 and for all odd primes (the only even prime number is the number 2) p.

In the two centuries following its conjecture (16371839), Fermat's Last Theorem was proven
for three odd prime exponents p = 3, 5 and 7. The case p = 3 was first stated by Abu-Mahmud
Khojandi (10th century), but his attempted proof of the theorem was incorrect.[50] In 1770,
Leonhard Euler gave a proof of p = 3,[51] but his proof by infinite descent[52] contained a major
gap.[53] However, since Euler himself had proven the lemma necessary to complete the proof in
other work, he is generally credited with the first proof.[54] Independent proofs were published[55]
by Kausler (1802),[25] Legendre (1823, 1830),[27][56] Calzolari (1855),[57] Gabriel Lam (1865),[58]
Peter Guthrie Tait (1872),[59] Gnther (1878),[60] Gambioli (1901),[36] Krey (1909),[61] Rychlk
(1910),[41] Stockhaus (1910),[62] Carmichael (1915),[63] Johannes van der Corput (1915),[64] Axel
Thue (1917),[65] and Duarte (1944).[66] The case p = 5 was proven[67] independently by Legendre
and Peter Dirichlet around 1825.[68] Alternative proofs were developed[69] by Carl Friedrich Gauss
(1875, posthumous),[70] Lebesgue (1843),[71] Lam (1847),[72] Gambioli (1901),[36][73] Werebrusow
(1905),[74] Rychlk (1910),[75] van der Corput (1915),[64] and Guy Terjanian (1987).[76] The case
p = 7 was proven[77] by Lam in 1839.[78] His rather complicated proof was simplified in 1840 by
Lebesgue,[79] and still simpler proofs[80] were published by Angelo Genocchi in 1864, 1874 and
1876.[81] Alternative proofs were developed by Thophile Ppin (1876)[82] and Edmond Maillet
(1897).[83]

Fermat's Last Theorem has also been proven for the exponents n = 6, 10, and 14. Proofs for n = 6
have been published by Kausler,[25] Thue,[84] Tafelmacher,[85] Lind,[86] Kapferer,[87] Swift,[88] and
Breusch.[89] Similarly, Dirichlet[90] and Terjanian[91] each proved the case n = 14, while Kapferer[87]
and Breusch[89] each proved the case n = 10. Strictly speaking, these proofs are unnecessary,
since these cases follow from the proofs for n = 3, 5, and 7, respectively. Nevertheless, the
reasoning of these even-exponent proofs differs from their odd-exponent counterparts. Dirichlet's
proof for n = 14 was published in 1832, before Lam's 1839 proof for n = 7.[92]

Many proofs for specific exponents use Fermat's technique of infinite descent, which Fermat
used to prove the case n = 4, but many do not. However, the details and auxiliary arguments are
often ad hoc and tied to the individual exponent under consideration.[93] Since they became ever
more complicated as p increased, it seemed unlikely that the general case of Fermat's Last
Theorem could be proven by building upon the proofs for individual exponents.[93] Although
some general results on Fermat's Last Theorem were published in the early 19th century by Niels
Henrik Abel and Peter Barlow,[94][95] the first significant work on the general theorem was done
by Sophie Germain.[96]

Sophie Germain
Main article: Sophie Germain
In the early 19th century, Sophie Germain developed several novel approaches to prove Fermat's
Last Theorem for all exponents.[97] First, she defined a set of auxiliary primes constructed from
the prime exponent p by the equation = 2hp+1, where h is any integer not divisible by three.
She showed that if no integers raised to the pth power were adjacent modulo (the non-
consecutivity condition), then must divide the product xyz. Her goal was to use mathematical
induction to prove that, for any given p, infinitely many auxiliary primes satisfied the non-
consecutivity condition and thus divided xyz; since the product xyz can have at most a finite
number of prime factors, such a proof would have established Fermat's Last Theorem. Although
she developed many techniques for establishing the non-consecutivity condition, she did not
succeed in her strategic goal. She also worked to set lower limits on the size of solutions to
Fermat's equation for a given exponent p, a modified version of which was published by Adrien-
Marie Legendre. As a byproduct of this latter work, she proved Sophie Germain's theorem,
which verified the first case of Fermat's Last Theorem (namely, the case in which p does not
divide xyz) for every odd prime exponent less than 100.[97][98] Germain tried unsuccessfully to
prove the first case of Fermat's Last Theorem for all even exponents, specifically for n = 2p,
which was proven by Guy Terjanian in 1977.[99] In 1985, Leonard Adleman, Roger Heath-Brown
and tienne Fouvry proved that the first case of Fermat's Last Theorem holds for infinitely many
odd primes p.[100]

Ernst Kummer and the theory of ideals


In 1847, Gabriel Lam outlined a proof of Fermat's Last Theorem based on factoring the
equation xp + yp = zp in complex numbers, specifically the cyclotomic field based on the roots of
the number 1. His proof failed, however, because it assumed incorrectly that such complex
numbers can be factored uniquely into primes, similar to integers. This gap was pointed out
immediately by Joseph Liouville, who later read a paper that demonstrated this failure of unique
factorisation, written by Ernst Kummer.

Kummer set himself the task of determining whether the cyclotomic field could be generalized to
include new prime numbers such that unique factorisation was restored. He succeeded in that
task by developing the ideal numbers. Using the general approach outlined by Lam, Kummer
proved both cases of Fermat's Last Theorem for all regular prime numbers. However, he could
not prove the theorem for the exceptional primes (irregular primes) which conjecturally occur
approximately 39% of the time; the only irregular primes below 100 are 37, 59 and 67.

Mordell conjecture
In the 1920s, Louis Mordell posed a conjecture that implied that Fermat's equation has at most a
finite number of nontrivial primitive integer solutions if the exponent n is greater than two.[101]
This conjecture was proven in 1983 by Gerd Faltings,[102] and is now known as Faltings' theorem.

Computational studies
In the latter half of the 20th century, computational methods were used to extend Kummer's
approach to the irregular primes. In 1954, Harry Vandiver used a SWAC computer to prove
Fermat's Last Theorem for all primes up to 2521.[103] By 1978, Samuel Wagstaff had extended
this to all primes less than 125,000.[104] By 1993, Fermat's Last Theorem had been proven for all
primes less than four million.[105]

Connection with elliptic curves


The ultimately successful strategy for proving Fermat's Last Theorem was by proving the
modularity theorem. The strategy was first described by Gerhard Frey in 1984.[106] Frey noted
that if Fermat's equation had a solution (a, b, c) for exponent p > 2, the corresponding elliptic
curve[note 2]

y2 = x (x ap)(x + bp)

would have such unusual properties that the curve would likely violate the modularity theorem.
[107]
This theorem, first conjectured in the mid-1950s and gradually refined through the 1960s,
states that every elliptic curve is modular, meaning that it can be associated with a unique
modular form.

Following this strategy, the proof of Fermat's Last Theorem required two steps. First, it was
necessary to show that Frey's intuition was correct: that the above elliptic curve, if it exists, is
always non-modular. Frey did not succeed in proving this rigorously; the missing piece was
identified by Jean-Pierre Serre. This missing piece, the so-called "epsilon conjecture", was
proven by Ken Ribet in 1986. Second, it was necessary to prove a special case of the modularity
theorem. This special case (for semistable elliptic curves) was proven by Andrew Wiles in 1995.

Thus, the epsilon conjecture showed that any solution to Fermat's equation could be used to
generate a non-modular semistable elliptic curve, whereas Wiles' proof showed that all such
elliptic curves must be modular. This contradiction implies that there can be no solutions to
Fermat's equation, thus proving Fermat's Last Theorem.

Wiles's general proof


British mathematician Andrew Wiles
Main article: Wiles's proof of Fermat's Last Theorem

Ribet's proof of the epsilon conjecture in 1986 accomplished the first half of Frey's strategy for
proving Fermat's Last Theorem. Upon hearing of Ribet's proof, Andrew Wiles decided to commit
himself to accomplishing the second half: proving a special case of the modularity theorem (then
known as the TaniyamaShimura conjecture) for semistable elliptic curves.[108] Wiles worked on
that task for six years in almost complete secrecy. He based his initial approach on his area of
expertise, Horizontal Iwasawa theory, but by the summer of 1991, this approach seemed
inadequate to the task.[109] In response, he exploited an Euler system recently developed by Victor
Kolyvagin and Matthias Flach. Since Wiles was unfamiliar with such methods, he asked his
Princeton colleague, Nick Katz, to check his reasoning over the spring semester of 1993.[110]

By mid-1993, Wiles was sufficiently confident of his results that he presented them in three
lectures delivered on June 2123, 1993 at the Isaac Newton Institute for Mathematical Sciences.
[111]
Specifically, Wiles presented his proof of the TaniyamaShimura conjecture for semistable
elliptic curves; together with Ribet's proof of the epsilon conjecture, this implied Fermat's Last
Theorem. However, it soon became apparent that Wiles's initial proof was incorrect. A critical
portion of the proof contained an error in a bound on the order of a particular group. The error
was caught by several mathematicians refereeing Wiles's manuscript including Katz,[112] who
alerted Wiles on 23 August 1993.[113]

Wiles and his former student Richard Taylor spent almost a year trying to repair the proof,
without success.[114] On 19 September 1994, Wiles had a flash of insight that the proof could be
saved by returning to his original Horizontal Iwasawa theory approach, which he had abandoned
in favour of the KolyvaginFlach approach.[115] On 24 October 1994, Wiles submitted two
manuscripts, "Modular elliptic curves and Fermat's Last Theorem"[116] and "Ring theoretic
properties of certain Hecke algebras",[117] the second of which was co-authored with Taylor. The
two papers were vetted and published as the entirety of the May 1995 issue of the Annals of
Mathematics. These papers established the modularity theorem for semistable elliptic curves, the
last step in proving Fermat's Last Theorem, 358 years after it was conjectured.

Exponents other than positive integers


Rational exponents

All solutions of the Diophantine equation when n=1 were computed by


[118] th
Lenstra in 1992. In the case in which the m roots are required to be real and positive, all
solutions are given by[119]

for positive integers r, s, t with s and t coprime.


In 2004, for n>2, Bennett, Glass, and Szekely proved that if gcd(n,m)=1, then there are integer
solutions if and only if 6 divides m, and , and are different complex 6th roots of
the same real number.[120]

Negative exponents

n = 1

All primitive (pairwise coprime) integer solutions to can be written as[121]

for positive, coprime integers m, n.

n = 2

The case n = 2 also has an infinitude of solutions, and these have a geometric interpretation in
terms of right triangles with integer sides and an integer altitude to the hypotenuse.[122][123] All
primitive solutions to are given by

for coprime integers u, v with v > u. The geometric interpretation is that a and b are the integer
legs of a right triangle and d is the integer altitude to the hypotenuse. Then the hypotenuse itself
is the integer

so (a, b, c) is a Pythagorean triple.

Integer n < 2

There are no solutions in integers for for integer n < 2. If there were, the
equation could be multiplied through by to obtain ,
which is impossible by Fermat's Last Theorem.

Did Fermat possess a general proof?


The mathematical techniques used in Fermat's "marvelous" proof are unknown. Only one
detailed proof of Fermat has survived, the above proof that no three coprime integers (x, y, z)
satisfy the equation x4 y4 = z2.

Taylor and Wiles's proof relies on mathematical techniques developed in the twentieth century,
which would be alien to mathematicians who had worked on Fermat's Last Theorem even a
century earlier. Fermat's alleged "marvellous proof", by comparison, would have had to be
elementary, given mathematical knowledge of the time, and so could not have been the same as
Wiles' proof. Most mathematicians and science historians doubt that Fermat had a valid proof of
his theorem for all exponents n.[citation needed]

Harvey Friedman's grand conjecture implies that Fermat's last theorem can be proved in
elementary arithmetic, a rather weak form of arithmetic with addition, multiplication,
exponentiation, and a limited form of induction for formulas with bounded quantifiers.[124] Any
such proof would be elementary but possibly too long to write down.

Monetary prizes
In 1816 and again in 1850, the French Academy of Sciences offered a prize for a general proof of
Fermat's Last Theorem.[125] In 1857, the Academy awarded 3000 francs and a gold medal to
Kummer for his research on ideal numbers, although he had not submitted an entry for the prize.
[126]
Another prize was offered in 1883 by the Academy of Brussels.[127]

In 1908, the German industrialist and amateur mathematician Paul Wolfskehl bequeathed
100,000 marks to the Gttingen Academy of Sciences to be offered as a prize for a complete
proof of Fermat's Last Theorem.[128] On 27 June 1908, the Academy published nine rules for
awarding the prize. Among other things, these rules required that the proof be published in a
peer-reviewed journal; the prize would not be awarded until two years after the publication; and
that no prize would be given after 13 September 2007, roughly a century after the competition
was begun.[129] Wiles collected the Wolfskehl prize money, then worth $50,000, on 27 June 1997.
[130]

Prior to Wiles' proof, thousands of incorrect proofs were submitted to the Wolfskehl committee,
amounting to roughly 10 feet (3 meters) of correspondence.[131] In the first year alone (1907
1908), 621 attempted proofs were submitted, although by the 1970s, the rate of submission had
decreased to roughly 34 attempted proofs per month. According to F. Schlichting, a Wolfskehl
reviewer, most of the proofs were based on elementary methods taught in schools, and often
submitted by "people with a technical education but a failed career".[132] In the words of
mathematical historian Howard Eves, "Fermat's Last Theorem has the peculiar distinction of
being the mathematical problem for which the greatest number of incorrect proofs have been
published."[127]

In popular culture
Main article: Fermat's Last Theorem in fiction
An episode in the television series Star Trek: The Next Generation, titled "The Royale",
refers to the theorem in the first act. Riker visits Captain Jean-Luc Picard in his ready
room to report only to find Picard puzzling over Fermat's last theorem. Picard's interest in
this theorem goes beyond the difficulty of the puzzle; he also feels humbled that despite
their advanced technology, they are still unable to solve a problem set forth by a man who
had no computer.[133] An episode in Star Trek: Deep Space 9, titled "Facets", refers to the
theorem as well. In a scene involving O'Brien, Tobin Dax mentions continuing work on
his own attempt to solve Fermat's last theorem.[134]

"The Proof" Nova (PBS) documentary about Andrew Wiles's proof of Fermat's Last
Theorem.

On August 17, 2011, a Google doodle was shown on the Google homepage, showing a
blackboard with the theorem on it. When hovered over, it displays the text "I have
discovered a truly marvelous proof of this theorem, which this doodle is too small to
contain." This is a reference to the note made by Fermat in the margins of Arithmetica. It
commemorated the 410th birth anniversary of de Fermat.[135]

In the book The Girl Who Played with Fire, main character Lisbeth Salander becomes
obsessed with the theorem in the opening chapters of the book. Her continuing effort to
come up with a proof on her own is a running sub-plot throughout the story, and is used
as a way to demonstrate her exceptional intelligence. In the end she comes up with a
proof (the actual proof is not featured in the book). But after being shot in the head and
surviving, she has lost the proof.

In the Harold Ramis re-make of the movie Bedazzled, starring Brendan Fraser and
Elizabeth Hurley, Fermat's Last Theorem appears written on the chalkboard in the
classroom that the protagonist Elliot finds himself teleported to after he aborts his failed
fourth wish. In the director's commentary for the DVD release, director Ramis comments
that nobody has seemed to notice that the equation on the board is Fermat's Last
Theorem.

In Doctor Who, Season 5 Episode 1 "The Eleventh Hour", the Doctor transmits a proof of
Fermat's Last Theorem by typing it in just a few seconds on Jeff's laptop to prove his
genius to a collection of world leaders discussing the latest threat to the human race. This
implies that the Doctor knew a proof which was quite short and easy for others to
comprehend.

In The IT Crowd, Series 3 Episode 6 "Calendar Geeks" Fermat's Last Theorem is


referenced during a photo shoot for a calendar about geeks and achievements in Science
and Mathematics.

The song "Bizarro Genius Baby" by MC Frontalot contains the lyrics "And no dust had
settled when shed disproved Fermat by finding A^3 + B^3 that = C^3".
Hypatia
From Wikipedia, the free encyclopedia
Jump to: navigation, search
For other uses see Hypatia (disambiguation)
Hypatia ()
ca. AD 351370
Born
Alexandria
AD 415
Died
Alexandria
Era Ancient philosophy
Region Alexandria, Egypt
School Neoplatonism
Main interests Mathematics, astronomy
Influenced by[show]
Influenced[show]

Hypatia (ca. AD 350370March 415) (pron.: /hape/ hy-PAY-sh; Ancient Greek: ;


Hypata) was a Greek Neoplatonist philosopher in Roman Egypt who was the first historically
noted woman in mathematics.[1] As head of the Platonist school at Alexandria, she also taught
philosophy and astronomy.[2][3][4][5][6]

As a Neoplatonist philosopher, she belonged to the mathematic tradition of the Academy of


Athens, as represented by Eudoxus of Cnidus;[7] she was of the intellectual school of the 3rd
century thinker Plotinus, which encouraged logic and mathematical study in place of empirical
enquiry and strongly encouraged law in place of nature.[1]

According to the only contemporary source, Hypatia was murdered by a Christian mob after
being accused of exacerbating a conflict between two prominent figures in Alexandria: the
governor Orestes and the Bishop of Alexandria.[8] Kathleen Wider proposes that the murder of
Hypatia marked the end of Classical antiquity,[9] and Stephen Greenblatt observes that her
murder "effectively marked the downfall of Alexandrian intellectual life".[10] On the other hand,
Maria Dzielska and Christian Wildberg note that Hellenistic philosophy continued to flourish in
the 5th and 6th centuries, and perhaps until the age of Justinian.[11]

Contents
1 Life

2 Death

o 2.1 Events leading to her murder

3 Works

4 Legacy

o 4.1 Late Antiquity to the Age of Reason

o 4.2 19th century

o 4.3 20th century

o 4.4 21st century

5 Notes

6 References

7 Further reading

8 External links

Life
Hypatia, by Charles William Mitchell (1885).

The mathematician and philosopher Hypatia of Alexandria was the daughter of the
mathematician Theon Alexandricus (ca. 335405).[12] She was educated at Athens and in Italy.
Around AD 400, she became head of the Platonist school at Alexandria,[13][14][15] where she
imparted the knowledge of Plato and Aristotle to any student; the pupils included pagans,
Christians, and foreigners.[1][16][17]

The contemporary 5th-century sources do identify Hypatia of Alexandria as a practitioner and


teacher of the philosophy of Plato and Plotinus, but, two hundred years later, the 7th-century
Egyptian Coptic bishop John of Niki identified her as a Hellenistic pagan and that "she was
devoted at all times to magic, astrolabes and instruments of music, and she beguiled many people
through her Satanic wiles".[18][19] Not all Christians were as hostile towards her as John of Nikiu
or the monks who killed her: some Christians even used Hypatia as symbolic of Virtue.[1]

The Byzantine Suda encyclopaedia reported that Hypatia was "the wife of Isidore the
Philosopher" (apparently Isidore of Alexandria);[16] however, Isidore of Alexandria was not born
until long after Hypatia's death, and no other philosopher of that name contemporary with
Hypatia is known.[20] The Suda also stated that "she remained a virgin" and that she rejected a
suitor with her menstrual rags, saying that they demonstrated "nothing beautiful" about carnal
desire, an example of a Christian source using Hypatia as a symbol of Virtue.[16][21][22]

Hypatia corresponded with former pupil Synesius of Cyrene, who was tutored by her in the
philosophical school of Platonism and later became bishop of Ptolemais in AD 410, an exponent
of the Christian Holy Trinity doctrine.[23] Together with the references by the pagan philosopher
Damascius, these are the extant records left by Hypatia's pupils at the Platonist school of
Alexandria.[24] The contemporary Christian historiographer Socrates Scholasticus described her
in Ecclesiastical History:
There was a woman at Alexandria named Hypatia, daughter of the philosopher Theo
of men. For all men on account of her extraordinary dignity and virtue admired her th

Socrates Scholasticus, Ecclesiastical History

Death
Events leading to her murder

Two widely cited, but divergent texts describe the feud between Orestes, the prefect (or
Governor) of Alexandria and Cyril, the Bishop of Alexandria. The feud and the city-wide anger it
provoked ultimately brought about the death of Hypatia.

One source, the Historia Ecclesiastica (or "Ecclesiastical History") was written by Socrates
Scholasticus (who was himself a Christian), some time shortly after Hypatia's death in AD 415.
Scholasticus gives a more complete, less biased account of the feud between Orestes and Cyril,
and the role Hypatia played in the feud that resulted in her death.

The other source, The Chronicle,[25] written by John of Nikiu in Egypt, around 650 AD,
demonizes Hypatia and Orestes directly, while validating all Christians involved in the events
Nikiu describes. The Chronicle, in being more biased on the matter of the historical feud, omits
several points of the narrative that are detailed in Scholasticuss account.

Ecclesiastical History, Socrates Scholasticus[26]

Orestes, the governor of Alexandria, and Cyril, the Bishop of Alexandria, found themselves in a
bitter feud in which Hypatia would come to be one of the main points of contention. The feud,
which took place in 415 AD, began over the matter of Jewish dancing exhibitions in Alexandria.
Since these exhibitions attracted large crowds and were commonly prone to civil disorder of
varying degrees, Orestes published an edict which outlined new regulations for such gatherings
and posted it in the city's theater. Soon after, crowds gathered to read the edict, angry over the
new regulations that had been imposed upon them. At one such gathering, Hierax, a Christian
and devout follower of Cyril, read the edict and applauded the new regulations, which many
people felt was an attempt to incite the crowd into sedition. In what Scholasticus suspected as
Orestes' "jealousy [of] the growing power of the bishops[which] encroached on the
jurisdiction of the authorities", Orestes immediately ordered Hierax to be seized and publicly
tortured in the theater.

Upon hearing of this, Cyril threatened the Jews of Alexandria with "the utmost severities" if
harassment of Christians was not ceased at once. In response, the Jews of Alexandria grew only
more furious over Cyril's threat, and in their anger they eventually resorted to violence against
the Christians. They plotted to flush the Christians out at night by running through the streets,
claiming that the Church of Alexander was on fire. When the Christians responded to what they
were led to believe was the burning down of their church, "the Jews immediately fell upon and
slew them", using rings to recognize one another in the dark, while killing everyone else in sight.
When the morning came, the Jews of Alexandria could not hide their guilt, and Cyril, along with
many of his followers, took to the citys synagogues in search of the perpetrators of the night's
massacre.

After Cyril found all of the Jews in Alexandria, he ordered them to be stripped of all their
possessions, banished them from Alexandria, and allowed the remaining citizens to pillage the
goods they left behind. With Cyril's banishment of the Jews, "Orestes [...] was filled with great
indignation at these transactions, and was excessively grieved that a city of such magnitude
should have been suddenly bereft of so large a portion of its population". Because of this, the
feud between Cyril and Orestes only grew stronger, and both men wrote to the emperor
regarding the situation. Eventually, Cyril attempted to reach out to Orestes through several peace
overtures, including attempted mediation and, when that failed, showed him the Gospels.
Nevertheless, Orestes remained unmoved by such gestures.

Meanwhile, approximately 500 monks, who resided in the mountains of Nitria, and were "of a
very fiery disposition", heard of the ongoing feud between the Governor and Bishop, and shortly
thereafter descended into Alexandria, armed and prepared to fight alongside Cyril. Upon their
arrival in Alexandria, the monks quickly intercepted Orestes' chariot in town and proceeded to
bombard and harass him, calling him a pagan idolater. In response to such allegations, Orestes
countered that he was actually a Christian, and had even been baptized by Atticus, the Bishop of
Constantinople. The monks paid little attention to Orestes claims of Christianity, and one of the
monks, by the name of Ammonius, struck Orestes in the head with a rock, which caused him to
bleed profusely. At this point, Orestes guards fled for fear of their lives, but a nearby crowd of
Alexandrians came to his aid, and Ammonius was subsequently secured and ordered to be
tortured for his actions. Upon excessive torture, Ammonius died. Following the death of
Ammonius, Cyril ordered that he henceforth be remembered as a martyr. Such a proclamation
did not sit well with "sober-minded" Christians, as Scholasticus pointed out, seeing that he
"suffered the punishment due to his rashness[not because] he would not deny Christ", and this
fact, according to Scholasticus, became more apparent to Cyril through general lack of
enthusiasm for Ammonius's case for martyrdom.

Scholasticus then introduces Hypatia, the female philosopher of Alexandria and woman who
would become a target of the Christian anger that grew over the feud. Daughter of Theon, and a
teacher trained in the philosophical schools of Plato and Plotinus, she was admired by most men
for her dignity and virtue. Of the anger she provoked among Christians, Scholasticus writes,
Hypatia ultimately fell "victim to the political jealousy which at the time prevailed" - Orestes
was known to seek her counsel, and a rumor spread among the Christian community of
Alexandria in which she was blamed for his unwillingness to reconcile with Cyril. Therefore, a
mob of Christians gathered, led by a reader (i.e. a minor cleric) named Peter whom Scholasticus
calls a fanatic. They kidnapped Hypatia on her way home and took her to the "Church called
Caesareum. They then completely stripped her, and then murdered her with tiles". Socrates
Scholasticus was hence interpreted as saying that, while she was still alive, Hypatia's flesh was
torn off using oyster shells (tiles; the Greek word is ostrakois, which literally means
"oystershells" but the word was also used for brick tiles on the roofs of houses). Afterward, the
men proceeded to mutilate her, and finally burn her limbs. When news broke of Hypatia's
murder, it provoked great public denouncement; not only against Cyril, but against the whole
Alexandrian Christian community. Scholasticus closes with a lament: "Surely nothing can be
farther from the spirit of Christianity than the allowance of massacres, fights, and transactions of
that sort".

Chronicle, John of Nikiu[27]

Bishop John of Nikiu, who lived several hundred years after the events he describes, writes
bitterly of Hypatia, claiming that "she beguiled many people through (her) Satanic wiles".
Orestes, who Nikiu writes, was himself a victim of Hypatia's demonic charm, regularly honored
her, and took to abandoning the Christian Church in order to follow her teachings more closely.
Moreover, Orestes himself persuaded others to leave the Church in favor of Hypatia's
philosophical teachings, and went as far as to host such "unbelievers" at his house.

One day, Orestes published an edict "regarding public exhibitions in the city of Alexandria", and
all citizens gathered to read Orestes's edict. Cyril, curious to see why the edict caused such an
uproar, sent Hierax, a "Christian possessing understanding and intelligence", who although
opposed to paganism, did as Cyril asked and went to learn the nature of Orestes's edict.
Meanwhile, the Jews who gathered in anger over the edict, believed that Hierax had only come
for the sake of provocation (which, according to Scholasticus's text, was Hierax's intent). Upon
this assumption, Orestes had Hierax punished for a crime for which "he was wholly guiltless".

For the punishment and torture of Hierax, as well as the death of several monks, including
Ammonius, Cyril grew increasingly furious with Orestes. (Here, Nikiu blatantly ignores the
assault of Orestes by the 500 monks, of which Ammonius played an active role in bringing about
his torture and death.) Cyril then warned the Jews against any further harm upon the Christians.
However, with the support of Orestes (which was in no way implied by Scholasticus), the Jews
felt confident in defying Cyril's authority, and so one night ran through the streets proclaiming:
"The church of the apostolic Athanasius (Alexander) is on fire: come to its succour, all ye
Christians". The Christians responded to the claims only to be slaughtered by the Jews in a
coordinated ambush.

The next morning, all remaining Christians of the town came to Cyril with news of the massacre,
after which Cyril marched with them to purge the Jews from Alexandria. In so doing, Cyril
allowed the pillaging of their possessions, and soon after purified all the synagogues in the city
and made them into Churches (Scholasticus makes no mention of "purifying" the Synagogues).
In the expulsion of the Jews, Orestes was unable to offer them any assistance.

Shortly thereafter, a group of Christians, under Peter the magistrate, went looking for Hypatia,
the "pagan woman who had beguiled the people of the city and the prefect through her
enchantments". They found her sitting in a chair, at which point they seized and brought her to
"the great church, named Caesarion", where they proceeded to rip the clothes off of her body.
Following this, they took to dragging her through the streets of Alexandria until she died. Once
she had died, they burned her remains. Nikiu's description of Hypatia's death also differs from
Scholasticus's interpretation. Following the death of Hypatia, Bishop Cyril was named "the new
Theophilus". With the death of Hypatia, Nikiu writes, the Christians had expelled the last
remnant of pagan idolatry.

Socrates Scholasticus (born after 380


John of Niki (7th century)
AD, died after 439 AD))

Yet even she fell a victim to the political And, in those days, there appeared in
jealousy which at that time prevailed. For Alexandria a female philosopher, a pagan
as she had frequent interviews with named Hypatia, and she was devoted at all
Orestes, it was calumniously reported times to magic, astrolabes, and instruments
among the Christian populace, that it was of music, and she beguiled many people
she who prevented Orestes from being through Satanic wiles . . . A multitude of
reconciled to the bishop. Some of them believers in God arose under the guidance
therefore, hurried away by a fierce and of Peter the Magistrate . . . and they
bigoted zeal, whose ringleader was a reader proceeded to seek for the pagan woman
named Peter, waylaid her returning home, who had beguiled the people of the city and
and dragging her from her carriage, they the Prefect through her enchantments. And
took her to the church called Caesareum, when they learnt the place where she was,
where they completely stripped her, and they proceeded to her and found her . . .
then murdered her with tiles. After tearing they dragged her along till they brought her
her body in pieces, they took her mangled to the great church, named Caesareum.
limbs to a place called Cinaron, and there Now this was in the days of the fast. And
burnt them.[28] they tore off her clothing and dragged her . .
. through the streets of the city till she died.
And they carried her to a place named
Cinaron, and they burned her body with
fire.[18]

Works
Cameron's 1867 photograph Hypatia

An actress, possibly Mary Anderson, in the title role of the play Hypatia, circa 1900.

No written work, widely recognized by scholars as Hypatia's own has survived to the present
time. Many of the works commonly attributed to her are believed to have been collaborative
works with her father, Theon Alexandricus, this kind of authorial uncertainty being typical for
female philosophers in Antiquity.[29]

A partial list of Hypatia's works as mentioned by other antique and medieval authors or as
posited by modern authors:

A commentary on the 13-volume Arithmetica by Diophantus.[16]

A commentary on the Conics of Apollonius.[16]


Edited the existing version of Ptolemy's Almagest.[30]

Edited her father's commentary on Euclid's Elements[31]

She wrote a text "The Astronomical Canon".[16] (Either a new edition of Ptolemy's Handy
Tables or the aforementioned Almagest.)[32][33]

Her contributions to science are reputed to include the charting of celestial bodies[6] and the
invention of the hydrometer,[34] used to determine the relative density (or specific gravity) of
liquids. However, the hydrometer was invented before Hypatia, and already known in her time.
[35][36]

Her student Synesius, bishop of Cyrene, wrote a letter describing his construction of an
astrolabe.[37] Earlier astrolabes predate that of Synesius by at least a century,[38][39] and Hypatia's
father had gained fame for his treatise on the subject.[40] However, Synesius claimed that his was
an improved model.[41] Synesius also sent Hypatia a letter describing a hydrometer, and
requesting her to have one constructed for him.[42]

Legacy
Late Antiquity to the Age of Reason

Shortly after her murder, there appeared under Hypatia's name a forged anti-Christian letter.[43]
The Neoplatonist historian Damascius (ca. AD 458538) was "anxious to exploit the scandal of
Hypatia's death", and attributed responsibility for her murder to Bishop Cyril and his Christian
followers; that historical account is contained in the Suda.[44] Damascius's account of the
Christian murder of Hypatia is the sole historical source attributing responsibility to Bishop
Cyril.[45] Maria Dzielska proposes that the bishop's body guards might have murdered Hypatia.[46]

The intellectual Eudokia Makrembolitissa (10211096), the second wife of Byzantine Emperor
Constantine X Doukas, was described by the historian Nicephorus Gregoras as a "second
Hypatia".[47]

Centuries later, the early 18th-century deist scholar John Toland used the murder of Hypatia as
the basis for the anti-Catholic tract Hypatia: Or the History of a most beautiful, most vertuous,
most learned, and every way accomplishd Lady; who was torn to pieces by the Clergy of
Alexandria, to gratify the pride, emulation, and cruelty of their Archbishop, commonly, but
undeservedly, stil'd St. Cyril.[48]

In turn, the Christians defended themselves from Toland with The History of Hypatia, a most
Impudent School-Mistress of Alexandria: Murder'd and torn to Pieces by the Populace, in
Defence of Saint Cyril and the Alexandrian Clergy from the Aspersions of Mr. Toland, by
Thomas Lewis[disambiguation needed], in 1721.[49]

19th century
In the 19th century, interest in the "literary legend of Hypatia" began to rise.[50] Diodata Saluzzo
Roero's 1827 Ipazia ovvero delle Filosofie suggested that Cyril had actually converted Hypatia to
Christianity, and that she had been killed by a "treacherous" priest.

In 1843, German authors Soldan and Heppe argued in their highly influential History of the
Witchcraft Trials that Hypatia may have been, in effect, the first famous "witch" punished under
Christian authority (see Witch-hunt).[51]

In his 1847 Hypatie and 1857 Hypatie et Cyrille, French poet Charles-Marie-Ren Leconte de
Lisle portrayed Hypatia as the epitome of "vulnerable truth and beauty".[52]

Charles Kingsley's 1853 novel Hypatia or New Foes with an Old Face, which portrayed the
scholar as a "helpless, pretentious, and erotic heroine",[53] recounted her conversion by a Jewish-
Christian named Raphael Aben-Ezra after supposedly becoming disillusioned with Orestes.

In 1867, the early photographer Julia Margaret Cameron created a portrait of the scholar as a
young woman.[54]

20th century

Some authors mention her in passing, such as Marcel Proust, who dropped her name in the last
sentence of "Madame Swann at Home," the first section of Within a Budding Grove.

Some characters are named after her, such as Hypatia Cade, a precocious child and main
character in the science fiction novel The Ship Who Searched by Mercedes Lackey and Anne
McCaffrey.

Rinne Groff's 2000 play The Five Hysterical Girls Theorem features a character named Hypatia
who lives silently, in fear that she will suffer the fate of her namesake.

Hypatia is the name of a 'shipmind' (ship computer) in The Boy Who Would Live Forever, a novel
in Frederik Pohl's Heechee series.

Umberto Eco's novel Baudolino sees the protagonist meet a secluded society of satyr-like
creatures who all take their name and philosophy from Hypatia.

A fictional version of the historic character appears in several works and indeed series, such as

The Heirs of Alexandria series written by Mercedes Lackey, Eric Flint and Dave Freer, is
an alternate history in which Hypatia was converted to Christianity by John Chrysostom,
which saved her life and enabled her to stop the mob from destroying the Library of
Alexandria; the books, taking place in the world of 1530 resulting from the above,
include copies from the alternate Hypatia's influential correspondence with John and St.
Augustine.
The Corto Maltese adventure Fable of Venice, by characteristic superposition of
anachronistic elements, sees Hypatia preside over an intellectual salon in pre-Fascist
Italy;

As a recurring character in Mark London Williams' juvenile fiction Danger Boy.

She also appears, briefly, as one of the kidnapped scientists and philosophers in the Doctor Who
episode Time and the Rani.

American astronomer Carl Sagan, in Cosmos: A Personal Voyage, gave a detailed speculative
description of Hypatia's death, linking it with the destruction of the Library of Alexandria.

A more scholarly historical study of her, Hypatia of Alexandria by Maria Dzielska (translated
into English by F. Lyra, published by Harvard University Press), was named by Choice
Magazine as an "Outstanding Academic Book of 1995, Philosophy Category".

She has been claimed by second wave feminism, most prominently as Hypatia: A Journal of
Feminist Philosophy, published since 1986 by Indiana University Press.

Judy Chicago's large-scale The Dinner Party awards her a place-setting, and other artistic works
draw on or are based on Hypatia.[citation needed]

A central character in Iain Pears' The Dream of Scipio is a woman philosopher clearly modeled
on (though not identical with) Hypatia.[citation needed]

The last two centuries have seen Hypatia's name honored in the sciences, especially astronomy.
238 Hypatia, a main belt asteroid discovered in 1884, was named for her. The lunar crater
Hypatia was named for her, in addition to craters named for her father Theon and for Cyril. The
180 km Rimae Hypatia is located north of the crater, one degree south of the equator, along the
Mare Tranquillitatis.[55]

By the end of the 20th century Hypatia's name was applied to projects ranging in scope from an
Adobe typeface (Hypatia Sans Pro),[56] to a cooperative community house in Madison,
Wisconsin. A genus of moth also bears her name.

21st century

Her life continues to be fictionalized by authors in many countries and languages. Two recent
examples are Ipazia, scienziata alessandrina by Adriano Petta (translated from the Italian in
2004 as Hypatia: Scientist of Alexandria), and Hypatia y la eternidad (Hypatia and Eternity) by
Ramon Gal, a fanciful alternate history, in Spanish (2009).[57]

The 2008 novel Azazl, by Egyptian Muslim author Dr. Ysuf Zaydan, tells the story of the
religious conflict of that time through the eyes of a monk, including a substantial section on
Hypatia;[58] Zaydan's book has been criticized by Christians in Egypt.[59]
Her life is portrayed in the Malayalam novel Francis Itty Cora (2009) by T. D Ramakrishnan.

Examples in English include

Remembering Hypatia: A Novel of Ancient Egypt by Brian Trent,;[60]

Flow Down Like Silver, Hypatia of Alexandria (2009) by Ki Longfellow, the second in a
trilogy of the divine feminine, the first being The Secret Magdalene;[61]

The Plot to Save Socrates (2006) by Paul Levinson and his sequel novelette "Unburning
Alexandria" (2008) - where Hypatia turns out to have been a time-traveler from 21st
Century America.

More factually, Hypatia of Alexandria: Mathematician and Martyr (2007) is a brief (113 page)
biography by Michael Deakin, with a focus on her mathematical research. Hypatia has been
considered a universal genius.[62]

The 2009 movie Agora, directed by Alejandro Amenbar, focuses on Hypatia's final years.
Hypatia, portrayed by actress Rachel Weisz, is seen investigating the heliocentric model of the
solar system proposed by Aristarchus of Samos, and even anticipating the elliptical orbits
discovered by Johannes Kepler 1200 years later.

Aryabhata

From Wikipedia, the free encyclopedia

Jump to: navigation, search

For other uses, see Aryabhata (disambiguation).

Aryabhata
Statue of Aryabhata on the grounds of IUCAA,
Pune. As there is no known information
regarding his appearance, any image of
Aryabhata originates from an artist's
conception.

476 CE
Born
Patliputra in Magadha

Died 550 CE

Era Gupta era

Region India

Main interests Mathematics, Astronomy

Major works ryabhaya, Arya-siddhanta

Aryabhata pronunciation (helpinfo) (IAST: ryabhaa, Sanskrit: ) or Aryabhata I[1][2]


(476550 CE)[3][4] was the first in the line of great mathematician-astronomers from the classical
age of Indian mathematics and Indian astronomy. His most famous works are the ryabhaya
(499 CE, when he was 23 years old)[5] and the Arya-siddhanta.

The works of Aryabhata dealt with mainly mathematics and astronomy. He also worked on the
approximation for pi.

Contents
1 Biography

o 1.1 Name

1.1.1 Time and place of birth

o 1.2 Education

o 1.3 Other hypotheses

2 Works

o 2.1 Aryabhatiya

3 Mathematics

o 3.1 Place value system and zero

o 3.2 Approximation of

o 3.3 Trigonometry

o 3.4 Indeterminate equations

o 3.5 Algebra

4 Astronomy

o 4.1 Motions of the solar system

o 4.2 Eclipses

o 4.3 Sidereal periods

o 4.4 Heliocentrism

5 Legacy

6 See also

7 References

o 7.1 Other references

8 External links
Biography

Name

While there is a tendency to misspell his name as "Aryabhatta" by analogy with other names
having the "bhatta" suffix, his name is properly spelled Aryabhata: every astronomical text spells
his name thus,[6] including Brahmagupta's references to him "in more than a hundred places by
name".[7] Furthermore, in most instances "Aryabhatta" does not fit the metre either.[6]

Time and place of birth


Aryabhata mentions in the Aryabhatiya that it was composed 3,630 years into the Kali Yuga,
when he was 23 years old. This corresponds to 499 CE, and implies that he was born in 476.[4]

Aryabhata was born in Bihar, India[6] However, early Buddhist texts describe Ashmaka as being
further south, in dakshinapath or the Deccan, while other texts describe the Ashmakas as having
fought Alexander.

Education

It is fairly certain that, at some point, he went to Kusumapura for advanced studies and lived
there for some time.[8] Both Hindu and Buddhist tradition, as well as Bhskara I (CE 629),
identify Kusumapura as Paliputra, modern Patna.[6] A verse mentions that Aryabhata was the
head of an institution (kulapati) at Kusumapura, and, because the university of Nalanda was in
Pataliputra at the time and had an astronomical observatory, it is speculated that Aryabhata might
have been the head of the Nalanda university as well.[6] Aryabhata is also reputed to have set up
an observatory at the Sun temple in Taregana, Bihar.[9]

Other hypotheses

Some archeological evidence suggests that Aryabhata could have originated from the present day
Kodungallur which was the historical capital city of Thiruvanchikkulam of ancient Kerala.[10] For
instance, one hypothesis was that amaka (Sanskrit for "stone") may be the region in Kerala that
is now known as Kouallr, based on the belief that it was earlier known as Koum-Kal-l-r
("city of hard stones"); however, old records show that the city was actually Koum-kol-r ("city
of strict governance"). Similarly, the fact that several commentaries on the Aryabhatiya have
come from Kerala were used to suggest that it was Aryabhata's main place of life and activity;
however, many commentaries have come from outside Kerala.

Aryabhata mentions "Lanka" on several occasions in the Aryabhatiya, but his "Lanka" is an
abstraction, standing for a point on the equator at the same longitude as his Ujjayini.[11]

Works
Aryabhata is the author of several treatises on mathematics and astronomy, some of which are
lost. His major work, Aryabhatiya, a compendium of mathematics and astronomy, was
extensively referred to in the Indian mathematical literature and has survived to modern times.
The mathematical part of the Aryabhatiya covers arithmetic, algebra, plane trigonometry, and
spherical trigonometry. It also contains continued fractions, quadratic equations, sums-of-power
series, and a table of sines.

The Arya-siddhanta, a lot work on astronomical computations, is known through the writings of
Aryabhata's contemporary, Varahamihira, and later mathematicians and commentators, including
Brahmagupta and Bhaskara I. This work appears to be based on the older Surya Siddhanta and
uses the midnight-day reckoning, as opposed to sunrise in Aryabhatiya. It also contained a
description of several astronomical instruments: the gnomon (shanku-yantra), a shadow
instrument (chhAyA-yantra), possibly angle-measuring devices, semicircular and circular
(dhanur-yantra / chakra-yantra), a cylindrical stick yasti-yantra, an umbrella-shaped device
called the chhatra-yantra, and water clocks of at least two types, bow-shaped and cylindrical.[12]

A third text, which may have survived in the Arabic translation, is Al ntf or Al-nanf. It claims that
it is a translation by Aryabhata, but the Sanskrit name of this work is not known. Probably dating
from the 9th century, it is mentioned by the Persian scholar and chronicler of India, Ab Rayhn
al-Brn.[12]

Aryabhatiya

Main article: Aryabhatiya

Direct details of Aryabhata's work are known only from the Aryabhatiya. The name
"Aryabhatiya" is due to later commentators. Aryabhata himself may not have given it a name.
His disciple Bhaskara I calls it Ashmakatantra (or the treatise from the Ashmaka). It is also
occasionally referred to as Arya-shatas-aShTa (literally, Aryabhata's 108), because there are 108
verses in the text. It is written in the very terse style typical of sutra literature, in which each line
is an aid to memory for a complex system. Thus, the explication of meaning is due to
commentators. The text consists of the 108 verses and 13 introductory verses, and is divided into
four pdas or chapters:

1. Gitikapada: (13 verses): large units of timekalpa, manvantra, and yuga


which present a cosmology different from earlier texts such as Lagadha's
Vedanga Jyotisha (c. 1st century BCE). There is also a table of sines (jya),
given in a single verse. The duration of the planetary revolutions during a
mahayuga is given as 4.32 million years.

2. Ganitapada (33 verses): covering mensuration (ketra vyvahra), arithmetic


and geometric progressions, gnomon / shadows (shanku-chhAyA), simple,
quadratic, simultaneous, and indeterminate equations
3. Kalakriyapada (25 verses): different units of time and a method for
determining the positions of planets for a given day, calculations concerning
the intercalary month (adhikamAsa), kShaya-tithis, and a seven-day week
with names for the days of week.

4. Golapada (50 verses): Geometric/trigonometric aspects of the celestial


sphere, features of the ecliptic, celestial equator, node, shape of the earth,
cause of day and night, rising of zodiacal signs on horizon, etc. In addition,
some versions cite a few colophons added at the end, extolling the virtues of
the work, etc.

The Aryabhatiya presented a number of innovations in mathematics and astronomy in verse


form, which were influential for many centuries. The extreme brevity of the text was elaborated
in commentaries by his disciple Bhaskara I (Bhashya, c. 600 CE) and by Nilakantha Somayaji in
his Aryabhatiya Bhasya, (1465 CE). He was not only the first to find the radius of the earth but
was the only one in ancient time including the Greeks and the Romans to find the volume of the
earth.

Mathematics

Place value system and zero

The place-value system, first seen in the 3rd century Bakhshali Manuscript, was clearly in place
in his work. While he did not use a symbol for zero, the French mathematician Georges Ifrah
explains that knowledge of zero was implicit in Aryabhata's place-value system as a place holder
for the powers of ten with null coefficients[13]

However, Aryabhata did not use the Brahmi numerals. Continuing the Sanskritic tradition from
Vedic times, he used letters of the alphabet to denote numbers, expressing quantities, such as the
table of sines in a mnemonic form.[14]

Approximation of

Aryabhata worked on the approximation for pi ( ), and may have come to the conclusion that
is irrational. In the second part of the Aryabhatiyam (gaitapda 10), he writes:

caturadhikam atamaaguam dvaistath sahasrm


ayutadvayavikambhasysanno vttapariha.
"Add four to 100, multiply by eight, and then add 62,000. By this rule the circumference of a
circle with a diameter of 20,000 can be approached." [15]

This implies that the ratio of the circumference to the diameter is ((4 + 100) 8 + 62000)/20000
= 62832/20000 = 3.1416, which is accurate to five significant figures.
It is speculated that Aryabhata used the word sanna (approaching), to mean that not only is this
an approximation but that the value is incommensurable (or irrational). If this is correct, it is
quite a sophisticated insight, because the irrationality of pi was proved in Europe only in 1761 by
Lambert.[16]

After Aryabhatiya was translated into Arabic (c. 820 CE) this approximation was mentioned in
Al-Khwarizmi's book on algebra.[12]

Trigonometry

In Ganitapada 6, Aryabhata gives the area of a triangle as

tribhujasya phalashariram samadalakoti bhujardhasamvargah

that translates to: "for a triangle, the result of a perpendicular with the half-side is the area."[17]

Aryabhata discussed the concept of sine in his work by the name of ardha-jya, which literally
means "half-chord". For simplicity, people started calling it jya. When Arabic writers translated
his works from Sanskrit into Arabic, they referred it as jiba. However, in Arabic writings, vowels
are omitted, and it was abbreviated as jb. Later writers substituted it with jaib, meaning "pocket"
or "fold (in a garment)". (In Arabic, jiba is a meaningless word.) Later in the 12th century, when
Gherardo of Cremona translated these writings from Arabic into Latin, he replaced the Arabic
jaib with its Latin counterpart, sinus, which means "cove" or "bay"; thence comes the English
since. Alphabetic code has been used by him to define a set of increments. If we use Aryabhata's
table and calculate the value of sin(30) (corresponding to hasjha) which is 1719/3438 = 0.5; the
value is correct. His alphabetic code is commonly known as the Aryabhata cipher.[18]

Indeterminate equations

A problem of great interest to Indian mathematicians since ancient times has been to find integer
solutions to equations that have the form ax + by = c, a topic that has come to be known as
diophantine equations. This is an example from Bhskara's commentary on Aryabhatiya:

Find the number which gives 5 as the remainder when divided by 8, 4 as the
remainder when divided by 9, and 1 as the remainder when divided by 7

That is, find N = 8x+5 = 9y+4 = 7z+1. It turns out that the smallest value for N is 85. In general,
diophantine equations, such as this, can be notoriously difficult. They were discussed extensively
in ancient Vedic text Sulba Sutras, whose more ancient parts might date to 800 BCE. Aryabhata's
method of solving such problems is called the kuaka ( ) method. Kuttaka means
"pulverizing" or "breaking into small pieces", and the method involves a recursive algorithm for
writing the original factors in smaller numbers. Today this algorithm, elaborated by Bhaskara in
621 CE, is the standard method for solving first-order diophantine equations and is often referred
to as the Aryabhata algorithm.[19] The diophantine equations are of interest in cryptology, and the
RSA Conference, 2006, focused on the kuttaka method and earlier work in the Sulbasutras.

Algebra

In Aryabhatiya Aryabhata provided elegant results for the summation of series of squares and
cubes:[20]

and

Astronomy

Aryabhata's system of astronomy was called the audAyaka system, in which days are reckoned
from uday, dawn at lanka or "equator". Some of his later writings on astronomy, which
apparently proposed a second model (or ardha-rAtrikA, midnight) are lost but can be partly
reconstructed from the discussion in Brahmagupta's khanDakhAdyaka. In some texts, he seems
to ascribe the apparent motions of the heavens to the Earth's rotation. He may have believed that
the planet's orbits as elliptical rather than circular.[21][22]

Motions of the solar system

Aryabhata correctly insisted that the earth rotates about its axis daily, and that the apparent
movement of the stars is a relative motion caused by the rotation of the earth, contrary to the
then-prevailing view in other parts of the world, that the sky rotated. This is indicated in the first
chapter of the Aryabhatiya, where he gives the number of rotations of the earth in a yuga,[23] and
made more explicit in his gola chapter:[24]

In the same way that someone in a boat going forward sees an unmoving [object]
going backward, so [someone] on the equator sees the unmoving stars going
uniformly westward. The cause of rising and setting [is that] the sphere of the stars
together with the planets [apparently?] turns due west at the equator, constantly
pushed by the cosmic wind.

Aryabhata described a geocentric model of the solar system, in which the Sun and Moon are
each carried by epicycles. They in turn revolve around the Earth. In this model, which is also
found in the Paitmahasiddhnta (c. CE 425), the motions of the planets are each governed by
two epicycles, a smaller manda (slow) and a larger ghra (fast). [25] The order of the planets in
terms of distance from earth is taken as: the Moon, Mercury, Venus, the Sun, Mars, Jupiter,
Saturn, and the asterisms."[12]
The positions and periods of the planets was calculated relative to uniformly moving points. In
the case of Mercury and Venus, they move around the Earth at the same mean speed as the Sun.
In the case of Mars, Jupiter, and Saturn, they move around the Earth at specific speeds,
representing each planet's motion through the zodiac. Most historians of astronomy consider that
this two-epicycle model reflects elements of pre-Ptolemaic Greek astronomy.[26] Another element
in Aryabhata's model, the ghrocca, the basic planetary period in relation to the Sun, is seen by
some historians as a sign of an underlying heliocentric model.[27]

Eclipses

Solar and lunar eclipses were scientifically explained by Aryabhata. He states that the Moon and
planets shine by reflected sunlight. Instead of the prevailing cosmogony in which eclipses were
caused by pseudo-planetary nodes Rahu and Ketu, he explains eclipses in terms of shadows cast
by and falling on Earth. Thus, the lunar eclipse occurs when the moon enters into the Earth's
shadow (verse gola.37). He discusses at length the size and extent of the Earth's shadow (verses
gola.3848) and then provides the computation and the size of the eclipsed part during an
eclipse. Later Indian astronomers improved on the calculations, but Aryabhata's methods
provided the core. His computational paradigm was so accurate that 18th century scientist
Guillaume Le Gentil, during a visit to Pondicherry, India, found the Indian computations of the
duration of the lunar eclipse of 30 August 1765 to be short by 41 seconds, whereas his charts (by
Tobias Mayer, 1752) were long by 68 seconds.[12]

Sidereal periods

Considered in modern English units of time, Aryabhata calculated the sidereal rotation (the
rotation of the earth referencing the fixed stars) as 23 hours, 56 minutes, and 4.1 seconds;[28] the
modern value is 23:56:4.091. Similarly, his value for the length of the sidereal year at 365 days,
6 hours, 12 minutes, and 30 seconds (365.25858 days)[29] is an error of 3 minutes and 20 seconds
over the length of a year (365.25636 days).[30]

Heliocentrism

As mentioned, Aryabhata advocated an astronomical model in which the Earth turns on its own
axis. His model also gave corrections (the gra anomaly) for the speeds of the planets in the sky
in terms of the mean speed of the sun. Thus, it has been suggested that Aryabhata's calculations
were based on an underlying heliocentric model, in which the planets orbit the Sun,[31][32][33]
though this has been rebutted.[34] It has also been suggested that aspects of Aryabhata's system
may have been derived from an earlier, likely pre-Ptolemaic Greek, heliocentric model of which
Indian astronomers were unaware,[35] though the evidence is scant.[36] The general consensus is
that a synodic anomaly (depending on the position of the sun) does not imply a physically
heliocentric orbit (such corrections being also present in late Babylonian astronomical texts), and
that Aryabhata's system was not explicitly heliocentric.[37]
Legacy

India's first satellite named after Aryabhata

Aryabhata's work was of great influence in the Indian astronomical tradition and influenced
several neighbouring cultures through translations. The Arabic translation during the Islamic
Golden Age (c. 820 CE), was particularly influenced. Some of his results are cited by Al-
Khwarizmi and in the 10th century Al-Biruni stated that Aryabhata's followers believed that the
Earth rotated on its axis.

His definitions of sine (jya), cosine (kojya), versine (utkrama-jya), and inverse sine (otkram jya)
influenced the birth of trigonometry. He was also the first to specify sine and versine (1 cos x)
tables, in 3.75 intervals from 0 to 90, to an accuracy of 4 decimal places.

In fact, modern names "sine" and "cosine" are mistranscriptions of the words jya and kojya as
introduced by Aryabhata. As mentioned, they were translated as jiba and kojiba in Arabic and
then misunderstood by Gerard of Cremona while translating an Arabic geometry text to Latin.
He assumed that jiba was the Arabic word jaib, which means "fold in a garment", L. sinus (c.
1150).[38]

Aryabhata's astronomical calculation methods were also very influential. Along with the
trigonometric tables, they came to be widely used in the Islamic world and used to compute
many Arabic astronomical tables (zijes). In particular, the astronomical tables in the work of the
Arabic Spain scientist Al-Zarqali (11th century) were translated into Latin as the Tables of
Toledo (12th c.) and remained the most accurate ephemeris used in Europe for centuries.

Calendric calculations devised by Aryabhata and his followers have been in continuous use in
India for the practical purposes of fixing the Panchangam (the Hindu calendar). In the Islamic
world, they formed the basis of the Jalali calendar introduced in 1073 CE by a group of
astronomers including Omar Khayyam,[39] versions of which (modified in 1925) are the national
calendars in use in Iran and Afghanistan today. The dates of the Jalali calendar are based on
actual solar transit, as in Aryabhata and earlier Siddhanta calendars. This type of calendar
requires an ephemeris for calculating dates. Although dates were difficult to compute, seasonal
errors were less in the Jalali calendar than in the Gregorian calendar.

India's first satellite Aryabhata and the lunar crater Aryabhata are named in his honour. An
Institute for conducting research in astronomy, astrophysics and atmospheric sciences is the
Aryabhatta Research Institute of Observational Sciences (ARIOS) near Nainital, India. The inter-
school Aryabhata Maths Competition is also named after him,[40] as is Bacillus aryabhata, a
species of bacteria discovered by ISRO scientists in 2009.[41]

ryabhaa's sine table


From Wikipedia, the free encyclopedia
Jump to: navigation, search

ryabhaa's sine table is a set of twenty-four of numbers given in the astronomical treatise
ryabhaiya composed by the fifth century Indian mathematician and astronomer ryabhaa
(476550 CE), for the computation of the half-chords of certain set of arcs of a circle. It is not a
table in the modern sense of a mathematical table; that is, it is not a set of numbers arranged into
rows and columns.[1][2]

Arc and chord of a circle


ryabhaa's table is also not a set of values of the trigonometric sine function in a conventional
sense; it is a table of the first differences of the values of trigonometric sines expressed in
arcminutes, and because of this the table is also referred to as ryabhaa's table of sine-
differences.[3][4]

ryabhaa's table was the first sine table ever constructed in the history of mathematics.[5] The
now lost tables of Hipparchus (c.190 BC c.120 BC) and Menelaus (c.70140 CE) and those of
Ptolemy (c.AD 90 c.168) were all tables of chords and not of half-chords.[5] ryabhaa's table
remained as the standard sine table of ancient India. There were continuous attempts to improve
the accuracy of this table. These endeavors culminated in the eventual discovery of the power
series expansions of the sine and cosine functions by Madhava of Sangamagrama (c.1350
c.1425), the founder of the Kerala school of astronomy and mathematics, and the tabulation of a
sine table by Madhava with values accurate to seven or eight decimal places.

Some historians of mathematics have argued that the sine table given in ryabhaiya was an
adaptation of earlier such tables constructed by mathematicians and astronomers of ancient
Greece.[6] David Pingree, one of America's foremost historians of the exact sciences in antiquity,
was an exponent of such a view. Assuming this hypothesis, G. J. Toomer[7][8][9] writes, "Hardly
any documentation exists for the earliest arrival of Greek astronomical models in India, or for
that matter what those models would have looked like. So it is very difficult to ascertain the
extent to which what has come down to us represents transmitted knowledge, and what is
original with Indian scientists. ... The truth is probably a tangled mixture of both."[10]

Contents
1 The table

o 1.1 The original table

o 1.2 In modern notations

2 ryabhaa's computational method

3 See also

4 References

The table
The original table

This article contains Indic text. Without proper


rendering support, you may see question marks
or boxes, misplaced vowels or missing conjuncts
instead of Indic text.

The stanza in ryabhaiya describing the sine table is reproduced below:

|
-- ||

In modern notations

The values encoded in ryabhaa's Sanskrit verse can be decoded using the numerical scheme
explained in Aryabhatiya, and the decoded numbers are listed in the table below. In the table, the
angle measures relevant to ryabhaa's sine table are listed in the second column. The third
column contains the list the numbers contained in the Sanskrit verse given above in Devanagari
script. For the convenience of users unable to read Devanagari, these word-numerals are
reproduced in the fourth column in ISO 15919 transliteration. The next column contains these
numbers in the Arabic numerals. ryabhaa's numbers are the first differences in the values of
sines. The corresponding value of sine (or more precisely, of jya) can be obtained by summing
up the differences up to that difference. Thus the value of jya corresponding to 18 45 is the sum
225 + 224 + 222 + 219 + 215 = 1105. For assessing the accuracy of ryabhaa's computations,
the modern values of jyas are given in the last column of the table.

In the Indian mathematical tradition, the sine ( or jya) of an angle is not a ratio of numbers. It is
the length of a certain line segment, a certain half-chord. The radius of the base circle is basic
parameter for the construction of such tables. Historically, several tables have been constructed
using different values for this parameter. ryabhaa has chosen the number 3438 as the value of
radius of the base circle for the computation of his sine table. The rationale of the choice of this
parameter is the idea of measuring the circumference of a circle in angle measures. In
astronomical computations distances are measured in degrees, minutes, seconds, etc. In this
measure, the circumference of a circle is 360 = (60 360) minutes = 21600 minutes. The radius
of the circle, the measure of whose circumference is 21600 minutes, is 21600 / 2 minutes.
Computing this using the value = 3.1416 known to Aryabhata one gets the radius of the circle
as 3438 minutes approximately. ryabhaa's sine table is based on this value for the radius of the
base circle. It has not yet been established who is the first ever to use this value for the base
radius. But Aryabhatiya is the earliest surviving text containing a reference to this basic constant.
[11]

Value in
Value in Modern
ryabhaa's ryabhaa'
Angle ( A ) ryabhaa's Value in value
Sl. numerical s
(in degrees, numerical notation Arabic of jya (A)
No notation value of
arcminutes) (in ISO 15919 numerals (3438 sin
(in jya (A)
transliteration) (A))
Devanagari)
1 03 45 makhi 225 225 224.8560
2 07 30 bhakhi 224 449 448.7490
3 11 15 phakhi 222 671 670.7205
4 15 00 dhakhi 219 890 889.8199
5 18 45 akhi 215 1105 1105.1089
6 22 30 akhi 210 1315 1315.6656
7 26 15 akhi 205 1520 1520.5885
8 30 00 hasjha 199 1719 1719.0000
9 33 45 skaki 191 1910 1910.0505
10 37 30 kiga 183 2093 2092.9218
11 41 15 ghaki 174 2267 2266.8309
12 45 00 kighva 164 2431 2431.0331
13 48 45 ghlaki 154 2585 2584.8253
14 52 30 kigra 143 2728 2727.5488
15 56 15 hakya 131 2859 2858.5925
16 60 00 dhaki 119 2978 2977.3953
17 63 45 kica 106 3084 3083.4485
18 67 30 sga 93 3177 3176.2978
19 71 15 jhaa 79 3256 3255.5458
20 75 00 va 65 3321 3320.8530
21 78 45 kla 51 3372 3371.9398
22 82 30 pta 37 3409 3408.5874
23 86 15 pha 22 3431 3430.6390
24 90 00 cha 7 3438 3438.0000

ryabhaa's computational method


The second section of ryabhaiya titled Ganitapda contains a stanza indicating a method for
the computation of the sine table. There are several ambiguities in correctly interpreting the
meaning of this verse. For example, the following is a translation of the verse given by Katz
wherein the words in square brackets are insertions of the translator and not translations of texts
in the verse.[11]

"When the second half-[chord] partitioned is less than the first half-chord, which is
[approximately equal to] the [corresponding] arc, by a certain amount, the remaining
[sine-differences] are less [than the previous ones] each by that amount of that divided by
the first half-chord."

Without any additional assumptions, no interpretation of this computational scheme has correctly
yielded all the numbers in the table compiled by ryabhaa in ryabhaiya.

You might also like