Jump to content

Fermat's Last Theorem

From Wikipedia, the free encyclopedia

This is an old revision of this page, as edited by 218.186.17.248 (talk) at 03:06, 12 November 2013 (→‎Computational studies). The present address (URL) is a permanent link to this revision, which may differ significantly from the current revision.

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 by Andrew Wiles it was in the Guinness Book of World Records for "most difficult mathematical problems".

Overview

The problem

Fermat's Last Theorem (known by this title historically although technically a conjecture, or unproven speculation, until proven in 1995) stood as an unsolved riddle in mathematics for over three centuries. The Theorem itself is a deceptively simple statement within mathematics that Fermat famously stated he had solved around 1637. His claim was discovered some 30 years later, after his death, as a bare statement in the margin of a book, but Fermat died without leaving any proof of his claim.

The claim eventually became one of the most famous unsolved problems of mathematics. The attempts made to prove it during that time prompted substantial development of number theory and over time Fermat's Last Theorem itself gained legendary prominence as an unsolved problem in popular mathematics. It is based upon the well known formula ("Pythagoras' Theorem") for a right-angle triangle discovered by the ancient Greek mathematician Pythagoras: a2 + b2 = c2.

The Pythagorean equation has an infinite number of whole-number solutions, representing the sides of a right-angle triangle; these solutions are known as Pythagorean triples. Fermat conjectured that the more general equation an + bn = cn had no solutions in positive integers a, b and c for any integer exponent greater than 2 — in other words that although a2 + b2 = c2 had an infinite number of whole-number solutions, the similar equations

a3 + b3 = c3
a4 + b4 = c4
an + bn = cn

for any other integer exponent n greater than 2 would have no solutions in positive integers. Although he claimed to have a general proof of his conjecture, Fermat left no details of his proof apart from the special case n = 4.

Subsequent developments and solution

With the special case n = 4 proven, the problem was to prove the theorem for exponents n that are prime numbers (this limitation is considered trivial to prove[note 1]). Over the next two centuries (1637–1839), the conjecture was proven for only the primes 3, 5, and 7, although Sophie Germain innovated and proved an approach which was relevant to an entire class of primes. In the mid-19th century, Ernst Kummer extended this and proved the theorem for all regular primes, leaving irregular primes to be analyzed individually. Building on Kummer's work and using sophisticated computer studies, other mathematicians were able to extend the proof to cover all prime exponents up to four million, but a proof for all exponents was inaccessible (meaning that mathematicians generally considered a proof to be either impossible, or at best exceedingly difficult, or not achievable with current knowledge).

The proof of Fermat's Last Theorem in full, for all n, was finally accomplished, however, after 358 years, by Andrew Wiles in 1995, an achievement for which he was honoured and received numerous awards. The solution came in a roundabout manner, from a completely different area of mathematics.

Around 1955 Japanese mathematicians Goro Shimura and Yutaka Taniyama suspected a link might exist between elliptic curves and modular forms, two completely different areas of mathematics. Known at the time as the Taniyama–Shimura-Weil conjecture, and (eventually) as the modularity theorem, it stood on its own, with no apparent connection to Fermat's Last Theorem. It was widely seen as significant and important in its own right, but was (like Fermat's equation) widely considered to be completely inaccessible to proof.

In 1984, Gerhard Frey noticed an apparent link between the modularity theorem and Fermat's Last Theorem. This potential link was confirmed two years later by Ken Ribet (see: Ribet's Theorem and Frey curve). On hearing this, English mathematician Andrew Wiles, who had a childhood fascination with Fermat's Last Theorem, decided to try to prove the modularity theorem as a way to prove Fermat's Last Theorem. In 1993, after six years working secretly on the problem, Wiles succeeded in proving enough of the modularity theorem to prove Fermat's Last Theorem. Wiles paper was massive in size and scope. A flaw was discovered in one part of his original paper during peer review and required a further year and collaboration with a past student, Richard Taylor, to resolve. As a result, the final proof in 1995 was accompanied by a second, smaller, joint paper to that effect. Wiles's achievement was reported widely in the popular press, and was popularized in books and television programs. The remaining parts of the modularity theorem were subsequently proven by other mathematicians, building on Wiles' work, between 1996 and 2001.

Mathematical history

Pythagoras and Diophantus

Pythagorean triples

A Pythagorean triple – named for the ancient Greek Pythagoras – is 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

Fermat's equation, xn + yn = zn with positive integer solutions, is an example of a Diophantine equation,[6] named for the 3rd-century Alexandrian mathematician, Diophantus, who studied them and developed methods for the solution of some kinds of Diophantine equations. 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.[7] Fermat's conjecture of his Last Theorem was inspired while reading a new edition of the Arithmetica,[8] which was translated into Latin and published in 1621 by Claude Bachet.[9]

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).[10] Solutions to linear Diophantine equations, such as 26x + 65y = 13, may be found using the Euclidean algorithm (c. 5th century BC).[11] 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 2]

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).[12]

Around 1637, Fermat wrote his Last Theorem in the margin of his copy of the Arithmetica next to Diophantus' sum-of-squares problem:[13]

Cubum autem in duos cubos, aut quadratoquadratum in duos quadratoquadratos, et generaliter nullam in infinitum ultra quadratum potestatem in duos eiusdem nominis fas est dividere cuius rei demonstrationem mirabilem sane detexi. Hanc marginis exiguitas non caperet. it is impossible to separate a cube into two cubes, or a fourth power into two fourth powers, or in general, any power higher than the second, into two like powers. I have discovered a truly marvelous proof of this, which this margin is too narrow to contain.[14]

It is not known whether Fermat had actually found a valid proof. His proof of one case (n = 4) by infinite descent has survived.[15] 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.[16] 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 Clément-Samuel Fermat produced a new edition of the book (1670) augmented with his father's comments.[17] The margin note became known as Fermat's Last Theorem,[18] as it was the last of Fermat's asserted theorems to remain unproven.[19]

Proofs 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.[20] 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 c4b4 = (a2)2.

Alternative proofs of the case n = 4 were developed later[21] by Frénicle de Bessy (1676),[22] Leonhard Euler (1738),[23] Kausler (1802),[24] Peter Barlow (1811),[25] Adrien-Marie Legendre (1830),[26] Schopis (1825),[27] Terquem (1846),[28] Joseph Bertrand (1851),[29] Victor Lebesgue (1853, 1859, 1862),[30] Theophile Pepin (1883),[31] Tafelmacher (1893),[32] David Hilbert (1897),[33] Bendz (1901),[34] Gambioli (1901),[35] Leopold Kronecker (1901),[36] Bang (1905),[37] Sommer (1907),[38] Bottari (1908),[39] Karel Rychlík (1910),[40] Nutzhorn (1912),[41] Robert Carmichael (1913),[42] Hancock (1931),[43] and Vrǎnceanu (1966).[44]

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),[45] Barbara (2007),[46] and Dolan (2011).[47]

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.[48] 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 (abc) 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 (adbdcd) 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 p (the only even prime number is the number 2).

In the two centuries following its conjecture (1637–1839), 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.[49] In 1770, Leonhard Euler gave a proof of p = 3,[50] but his proof by infinite descent[51] contained a major gap.[52] However, since Euler himself had proven the lemma necessary to complete the proof in other work, he is generally credited with the first proof.[53] Independent proofs were published[54] by Kausler (1802),[24] Legendre (1823, 1830),[26][55] Calzolari (1855),[56] Gabriel Lamé (1865),[57] Peter Guthrie Tait (1872),[58] Günther (1878),[59] Gambioli (1901),[35] Krey (1909),[60] Rychlík (1910),[40] Stockhaus (1910),[61] Carmichael (1915),[62] Johannes van der Corput (1915),[63] Axel Thue (1917),[64] and Duarte (1944).[65] The case p = 5 was proven[66] independently by Legendre and Peter Gustav Lejeune Dirichlet around 1825.[67] Alternative proofs were developed[68] by Carl Friedrich Gauss (1875, posthumous),[69] Lebesgue (1843),[70] Lamé (1847),[71] Gambioli (1901),[35][72] Werebrusow (1905),[73] Rychlík (1910),[74] van der Corput (1915),[63] and Guy Terjanian (1987).[75] The case p = 7 was proven[76] by Lamé in 1839.[77] His rather complicated proof was simplified in 1840 by Lebesgue,[78] and still simpler proofs[79] were published by Angelo Genocchi in 1864, 1874 and 1876.[80] Alternative proofs were developed by Théophile Pépin (1876)[81] and Edmond Maillet (1897).[82]

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,[24] Thue,[83] Tafelmacher,[84] Lind,[85] Kapferer,[86] Swift,[87] and Breusch.[88] Similarly, Dirichlet[89] and Terjanian[90] each proved the case n = 14, while Kapferer[86] and Breusch[88] 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 Lamé's 1839 proof for n = 7.[91]

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.[92] 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.[92] Although some general results on Fermat's Last Theorem were published in the early 19th century by Niels Henrik Abel and Peter Barlow,[93][94] the first significant work on the general theorem was done by Sophie Germain.[95]

Sophie Germain

In the early 19th century, Sophie Germain developed several novel approaches to prove Fermat's Last Theorem for all exponents.[96] First, she defined a set of auxiliary primes θ constructed from the prime exponent p by the equation θ = 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 θ (the non-consecutivity condition), then θ must divide the product xyz. Her goal was to use mathematical induction to prove that, for any given p, infinitely many auxiliary primes θ 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.[96][97] 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.[98] In 1985, Leonard Adleman, Roger Heath-Brown and Étienne Fouvry proved that the first case of Fermat's Last Theorem holds for infinitely many odd primes p.[99]

Ernst Kummer and the theory of ideals

In 1847, Gabriel Lamé 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 Lamé, 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.[100] This conjecture was proven in 1983 by Gerd Faltings,[101] and is now known as Faltings' theorem.

Computational studies

In the later 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.[102] By 1978, Samuel Wagstaff had extended this to all primes less than 125,000.[103] By 1993, Fermat's Last Theorem had been proven for all primes less than four million.[104]

However despite these efforts and their results, no proof existed of Fermat's Last Theorem. Proofs of individual exponents by their nature could never prove the general case: even if all exponents were verified up to an extremely large number X, a higher exponent beyond X might still exist for which the claim was not true. (This had been the case with some other past conjectures, and it could not be ruled out in this conjecture.)

Connection with elliptic curves

The strategy that ultimately led to a successful proof of Fermat's Last Theorem arose from the "astounding"[105]: 211  Taniyama–Shimura-Weil conjecture, proposed around 1955, which many mathematicians believed would be near to impossible to prove,[105]: 223  and which was linked in the 1980s by Gerhard Frey, Jean-Pierre Serre and Ken Ribet to Fermat's equation. By accomplishing a partial proof of this conjecture in 1995, Andrew Wiles ultimately succeeded in proving Fermat's Last Theorem, as well as leading the way to a full proof by others of what is now the modularity theorem.

The Taniyama–Shimura-Weil conjecture

Around 1955, Japanese mathematicians Goro Shimura and Yutaka Taniyama observed a possible link between two apparently completely distinct, branches of mathematics, elliptic curves and modular forms. The resulting modularity theorem (at the time known as the Taniyama–Shimura conjecture) states that every elliptic curve is modular, meaning that it can be associated with a unique modular form. It was initially dismissed as unlikely or highly speculative, and was taken more seriously when number theorist André Weil found evidence supporting it, but no proof; as a result the "astounding"[105]: 211  conjecture was often known as the Taniyama–Shimura-Weil conjecture. It became a part of the Langlands programme, a list of important conjectures needing proof or disproof.[105]: 211–215 

Even after gaining serious attention, the conjecture was seen by contemporary mathematicians as extraordinarily difficult or perhaps inaccessible to proof.[105]: 203–205, 223, 226  For example, Wiles' ex-supervisor John Coates states that it seemed "impossible to actually prove",[105]: 226  and Ken Ribet considered himself "one of the vast majority of people who believed [it] was completely inaccessible", adding that "Andrew Wiles was probably one of the few people on earth who had the audacity to dream that you can actually go and prove [it]." [105]: 223 

Frey's equation / Ribet's theorem

In 1984, Gerhard Frey noted a link between Fermat's equation and the modularity theorem, then still a conjecture. If Fermat's equation had any solution (a, b, c) for exponent p > 2, then it could be shown that the elliptic curve (now known as a Frey curve [note 3])

y2 = x (x − ap)(x + bp)

would have such unusual properties that it was unlikely to be modular.[106] This would conflict with the modularity theorem, which asserted that all elliptic curves are modular. As such, Frey observed that a proof of the Taniyama–Shimura-Weil conjecture would simultaneously prove Fermat's Last Theorem.[107] (Equally, a disproof or refutation of Fermat's Last Theorem would disprove the conjecture)

Following this strategy, a proof of Fermat's Last Theorem required two steps. First, it was necessary to show that Frey's intuition was correct: that if an elliptic curve were constructed in this way, using a set of numbers that were a solution of Fermat's equation, the resulting elliptic curve could not be modular. Frey did not quite succeed in proving this rigorously; the missing piece (the so-called "epsilon conjecture", now known as Ribet's theorem) was identified by Jean-Pierre Serre[citation needed] and proven in 1986 by Ken Ribet. Second, it was necessary to prove the modularity theorem – or at least to prove it for the sub-class of cases (known as semistable elliptic curves) which included Frey's equation – and this was widely believed inaccessible to proof by contemporary mathematicians.[105]: 203–205, 223, 226 

  • The modularity theorem – if proven – would mean all elliptic curves (or at least all semistable elliptic curves) are of necessity modular.
  • Ribet's theorem – proven in 1986 – showed that if a solution to Fermat's equation existed, it could be used to create a semistable elliptic curve that was not modular;
  • The contradiction would imply (if the modularity theorem were correct) that no solutions can exist to Fermat's equation – therefore proving Fermat's Last Theorem.

Wiles's general proof

British mathematician Andrew Wiles

Ribet's proof of the epsilon conjecture in 1986 accomplished the first of the two goals proposed by Frey. Upon hearing of Ribet's success, Andrew Wiles, an English mathematician with a childhood fascination with Fermat's Last Theorem, and a prior study area of elliptical equations, decided to commit himself to accomplishing the second half: proving a special case of the modularity theorem (then known as the Taniyama–Shimura conjecture) for semistable elliptic curves.[108]

Wiles worked on that task for six years in near-total secrecy, covering up his efforts by releasing prior work in small segments as separate papers and confiding only in his wife.[105]: 229–230  His initial study suggested proof by induction,[105]: 230–232, 249–252  and he based his initial work and first significant breakthrough on Galois theory[105]: 251–253, 259  before switching to an attempt to extend Horizontal Iwasawa theory for the inductive argument around 1990–91 when it seemed that there was no existing approach adequate to the problem.[105]: 258–259  However, by the summer of 1991, Iwasawa theory also seemed to not be reaching the central issues in the problem.[105]: 259–260 [109] In response, he approached colleagues to seek out any hints of cutting edge research and new techniques, and discovered an Euler system recently developed by Victor Kolyvagin and Matthias Flach which seemed "tailor made" for the inductive part of his proof.[105]: 260–261  Wiles studied and extended this approach, which worked. Since his work relied extensively on this approach, which was new to mathematics and to Wiles, in January 1993 he asked his Princeton colleague, Nick Katz, to check his reasoning for subtle errors. Their conclusion at the time was that the techniques used by Wiles seemed to be working correctly.[105]: 261–265 [110]

By mid-May 1993 Wiles felt able to tell his wife he thought he had solved the proof of Fermat's Last Theorem,[105]: 265  and by June he felt sufficiently confident to present his results in three lectures delivered on 21–23 June 1993 at the Isaac Newton Institute for Mathematical Sciences.[111] Specifically, Wiles presented his proof of the Taniyama–Shimura conjecture for semistable elliptic curves; together with Ribet's proof of the epsilon conjecture, this implied Fermat's Last Theorem. However, it became apparent during peer review that a critical point in the proof was incorrect. It 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 (in his role as reviewer),[112] who alerted Wiles on 23 August 1993.[113]

The error would not have rendered his work worthless – each part of Wiles' work was highly significant and innovative by itself, as were the many developments and techniques he had created in the course of his work, and only one part was affected.[105]: 289, 296–297  However without this part proven, there was no actual proof of Fermat's Last Theorem. Wiles spent almost a year trying to repair his proof, initially by himself and then in collaboration with Richard Taylor, without success.[114]

On 19 September 1994, on the verge of giving up, 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 Kolyvagin–Flach approach, this time strengthening it with expertise gained in Kolyvagin–Flach's 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 and proved that certain conditions were met which were needed to justify the corrected step in the main paper. 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.

Subsequent developments

The full Taniyama–Shimura–Weil conjecture was finally proved by Diamond (1996), Conrad, Diamond & Taylor (1999), and Breuil et al. (2001) who, building on Wiles' work, incrementally chipped away at the remaining cases until the full result was proved. The now fully proved conjecture became known as the modularity theorem.

Several other theorems in number theory similar to Fermat's Last Theorem also follow from the same reasoning, using the modularity theorem. For example: no cube can be written as a sum of two coprime n-th powers, n ≥ 3. (The case n = 3 was already known by Euler.)

Exponents other than positive integers

Rational exponents

All solutions of the Diophantine equation when n=1 were computed by Lenstra in 1992.[118] In the case in which the mth 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 x4y4 = z2. Note that this proves Fermat's Last Theorem for the case n=4, since the equation a4 + b4 = c4 can be written as c4b4 = (a2)2.

Taylor and Wiles's proof relies on mathematical techniques developed in the 20th century,[124] which would be unknown 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 any provable theorem (including Fermat's last theorem) can be proved using only elementary arithmetic, a rather weak form of arithmetic with addition, multiplication, exponentiation, and a limited form of induction for formulas with bounded quantifiers.[125] Any such proof would indeed be 'elementary' but could involve implausibly long proofs of millions—or millions of millions—of steps, and might be far too long to be Fermat's proof.

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.[126] 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.[127] Another prize was offered in 1883 by the Academy of Brussels.[128]

In 1908, the German industrialist and amateur mathematician Paul Wolfskehl bequeathed 100,000 marks to the Göttingen Academy of Sciences to be offered as a prize for a complete proof of Fermat's Last Theorem.[129] 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.[130] Wiles collected the Wolfskehl prize money, then worth $50,000, on 27 June 1997.[131]

Prior to Wiles' proof, thousands of incorrect proofs were submitted to the Wolfskehl committee, amounting to roughly 10 feet (3 meters) of correspondence.[132] In the first year alone (1907–1908), 621 attempted proofs were submitted, although by the 1970s, the rate of submission had decreased to roughly 3–4 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".[133] 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."[128]

See also

Footnotes

  1. ^ If the exponent "n" were not prime then it would be possible to write n as a product of two smaller integers (n = P*Q) in which P is a prime number, and then an = aP*Q = (aQ)P for each of a, b, and c, i.e. an equivalent solution would also have to exist for the prime power P which is smaller than N, as well.
  2. ^ For example,
  3. ^ This elliptic curve was first suggested in the 1960s by Yves Hellegouarch, but he did not call attention to its non-modularity. For more details, see Hellegouarch, Yves (2001). Invitation to the Mathematics of Fermat-Wiles. Academic Press. ISBN 978-0-12-339251-0.

References

  1. ^ Stark, pp. 151–155.
  2. ^ Stillwell J (2003). Elements of Number Theory. New York: Springer-Verlag. pp. 110–112. ISBN 0-387-95587-9.
  3. ^ Aczel, pp. 13–15
  4. ^ Singh, pp. 18–20.
  5. ^ Singh, p. 6.
  6. ^ Stark, pp. 145–146.
  7. ^ Singh, pp. 50–51.
  8. ^ Stark, p. 145.
  9. ^ Aczel, pp. 44–45; Singh, pp. 56–58.
  10. ^ Aczel, pp. 14–15.
  11. ^ Stark, pp. 44–47.
  12. ^ Friberg, pp. 333– 334.
  13. ^ Dickson, p. 731; Singh, pp. 60–62; Aczel, p. 9.
  14. ^ Panchishkin, p. 341
  15. ^ Dickson, pp. 615–616; Aczel, p. 44.
  16. ^ Ribenboim, pp. 13, 24.
  17. ^ Singh, pp. 62–66.
  18. ^ Dickson, p. 731.
  19. ^ Singh, p. 67; Aczel, p. 10.
  20. ^ Freeman L. "Fermat's One Proof". Retrieved 23 May 2009.
  21. ^ Ribenboim, pp. 15–24.
  22. ^ Frénicle de Bessy, Traité des Triangles Rectangles en Nombres, vol. I, 1676, Paris. Reprinted in Mém. Acad. Roy. Sci., 5, 1666–1699 (1729).
  23. ^ Euler L (1738). "Theorematum quorundam arithmeticorum demonstrationes". Comm. Acad. Sci. Petrop. 10: 125–146.. Reprinted Opera omnia, ser. I, "Commentationes Arithmeticae", vol. I, pp. 38–58, Leipzig:Teubner (1915).
  24. ^ a b c Kausler CF (1802). "Nova demonstratio theorematis nec summam, nec differentiam duorum cuborum cubum esse posse". Novi Acta Acad. Petrop. 13: 245–253.
  25. ^ Barlow P (1811). An Elementary Investigation of Theory of Numbers. St. Paul's Church-Yard, London: J. Johnson. pp. 144–145.
  26. ^ a b Legendre AM (1830). Théorie des Nombres (Volume II) (3rd ed.). Paris: Firmin Didot Frères. Reprinted in 1955 by A. Blanchard (Paris).
  27. ^ Schopis (1825). Einige Sätze aus der unbestimmten Analytik. Gummbinnen: Programm.
  28. ^ Terquem O (1846). "Théorèmes sur les puissances des nombres". Nouv. Ann. Math. 5: 70–87.
  29. ^ Bertrand J (1851). Traité Élémentaire d'Algèbre. Paris: Hachette. pp. 217–230, 395.
  30. ^ Lebesgue VA (1853). "Résolution des équations biquadratiques z2 = x4 ± 2my4, z2 = 2mx4y4, 2mz2 = x4 ± y4". J. Math. Pures Appl. 18: 73–86.
    Lebesgue VA (1859). Exercices d'Analyse Numérique. Paris: Leiber et Faraguet. pp. 83–84, 89.
    Lebesgue VA (1862). Introduction à la Théorie des Nombres. Paris: Mallet-Bachelier. pp. 71–73.
  31. ^ Pepin T (1883). "Étude sur l'équation indéterminée ax4 + by4 = cz2". Atti Accad. Naz. Lincei. 36: 34–70.
  32. ^ Tafelmacher WLA (1893). "Sobre la ecuación x4 + y4 = z4". Ann. Univ. Chile. 84: 307–320.
  33. ^ Hilbert D (1897). "Die Theorie der algebraischen Zahlkörper". Jahresbericht der Deutschen Mathematiker-Vereinigung. 4: 175–546. Reprinted in 1965 in Gesammelte Abhandlungen, vol. I by New York:Chelsea.
  34. ^ Bendz TR (1901). Öfver diophantiska ekvationen xn + yn = zn. Uppsala: Almqvist & Wiksells Boktrycken.
  35. ^ a b c Gambioli D (1901). "Memoria bibliographica sull'ultimo teorema di Fermat". Period. Mat. 16: 145–192.
  36. ^ Kronecker L (1901). Vorlesungen über Zahlentheorie, vol. I. Leipzig: Teubner. pp. 35–38. Reprinted by New York:Springer-Verlag in 1978.
  37. ^ Bang A (1905). "Nyt Bevis for at Ligningen x4y4 = z4, ikke kan have rationale Løsinger". Nyt Tidsskrift Mat. 16B: 35–36.
  38. ^ Sommer J (1907). Vorlesungen über Zahlentheorie. Leipzig: Teubner.
  39. ^ Bottari A (1908). "Soluzione intere dell'equazione pitagorica e applicazione alla dimostrazione di alcune teoremi dellla teoria dei numeri". Period. Mat. 23: 104–110.
  40. ^ a b Rychlik K (1910). "On Fermat's last theorem for n = 4 and n = 3 (in Bohemian)". Časopis Pěst. Mat. 39: 65–86.
  41. ^ Nutzhorn F (1912). "Den ubestemte Ligning x4 + y4 = z4". Nyt Tidsskrift Mat. 23B: 33–38.
  42. ^ Carmichael RD (1913). "On the impossibility of certain Diophantine equations and systems of equations". Amer. Math. Monthly. 20 (7). Mathematical Association of America: 213–221. doi:10.2307/2974106. JSTOR 2974106.
  43. ^ Hancock H (1931). Foundations of the Theory of Algebraic Numbers, vol. I. New York: Macmillan.
  44. ^ Vrǎnceanu G (1966). "Asupra teorema lui Fermat pentru n=4". Gaz. Mat. Ser. A. 71: 334–335. Reprinted in 1977 in Opera matematica, vol. 4, pp. 202–205, Bucureşti:Edit. Acad. Rep. Soc. Romana.
  45. ^ Grant, Mike, and Perella, Malcolm, "Descending to the irrational", Mathematical Gazette 83, July 1999, pp.263–267.
  46. ^ Barbara, Roy, "Fermat's last theorem in the case n=4", Mathematical Gazette 91, July 2007, 260–262.
  47. ^ Dolan, Stan, "Fermat's method of descente infinie", Mathematical Gazette 95, July 2011, 269–271.
  48. ^ Ribenboim, pp. 1–2.
  49. ^ Dickson, p. 545.
    O'Connor, John J.; Robertson, Edmund F., "Abu Mahmud Hamid ibn al-Khidr Al-Khujandi", MacTutor History of Mathematics Archive, University of St Andrews
  50. ^ Euler L (1770) Vollständige Anleitung zur Algebra, Roy. Acad. Sci., St. Petersburg.
  51. ^ Freeman L. "Fermat's Last Theorem: Proof for n = 3". Retrieved 23 May 2009.
  52. ^ Ribenboim, pp. 24–25; Mordell, pp. 6–8; Edwards, pp. 39–40.
  53. ^ Aczel, p. 44; Edwards, pp. 40, 52–54.
    J. J. Mačys (2007). "On Euler's hypothetical proof". Mathematical Notes. 82 (3–4): 352–356. doi:10.1134/S0001434607090088. MR 2364600.
  54. ^ Ribenboim, pp. 33, 37–41.
  55. ^ Legendre AM (1823). "Recherches sur quelques objets d'analyse indéterminée, et particulièrement sur le théorème de Fermat". Mém. Acad. Roy. Sci. Institut France. 6: 1–60. Reprinted in 1825 as the "Second Supplément" for a printing of the 2nd edition of Essai sur la Théorie des Nombres, Courcier (Paris). Also reprinted in 1909 in Sphinx-Oedipe, 4, 97–128.
  56. ^ Calzolari L (1855). Tentativo per dimostrare il teorema di Fermat sull'equazione indeterminata xn + yn = zn. Ferrara.
  57. ^ Lamé G (1865). "Étude des binômes cubiques x3 ± y3". C. R. Acad. Sci. Paris. 61: 921–924, 961–965.
  58. ^ Tait PG (1872). "Mathematical Notes". Proc. Roy. Soc. Edinburgh. 7: 144.
  59. ^ Günther S (1878). "Über die unbestimmte Gleichung x3 + y3 = z3". Sitzungsberichte Böhm. Ges. Wiss.: 112–120.
  60. ^ Krey H (1909). "Neuer Beweis eines arithmetischen Satzes". Math. Naturwiss. Blätter. 6: 179–180.
  61. ^ Stockhaus H (1910). Beitrag zum Beweis des Fermatschen Satzes. Leipzig: Brandstetter.
  62. ^ Carmichael RD (1915). Diophantine Analysis. New York: Wiley.
  63. ^ a b van der Corput JG (1915). "Quelques formes quadratiques et quelques équations indéterminées". Nieuw Archief Wisk. 11: 45–75.
  64. ^ Thue A (1917). "Et bevis for at ligningen A3 + B3 = C3 er unmulig i hele tal fra nul forskjellige tal A, B og C". Arch. Mat. Naturv. 34 (15). Reprinted in Selected Mathematical Papers (1977), Oslo:Universitetsforlaget, pp. 555–559.
  65. ^ Duarte FJ (1944). "Sobre la ecuación x3 + y3 + z3 = 0". Ciencias Fis. Mat. Naturales (Caracas). 8: 971–979.
  66. ^ Freeman L. "Fermat's Last Theorem: Proof for n = 5". Retrieved 23 May 2009.
  67. ^ Ribenboim, p. 49; Mordell, p. 8–9; Aczel, p. 44; Singh, p. 106.
  68. ^ Ribenboim, pp. 55–57.
  69. ^ Gauss CF (1875, posthumous). "Neue Theorie der Zerlegung der Cuben". Zur Theorie der complexen Zahlen, Werke, vol. II (2nd ed.). Königl. Ges. Wiss. Göttingen. pp. 387–391. {{cite book}}: Check date values in: |year= (help)
  70. ^ Lebesgue VA (1843). "Théorèmes nouveaux sur l'équation indéterminée x5 + y5 = az5". J. Math. Pures Appl. 8: 49–70.
  71. ^ Lamé G (1847). "Mémoire sur la résolution en nombres complexes de l'équation A5 + B5 + C5 = 0". J. Math. Pures Appl. 12: 137–171.
  72. ^ Gambioli D (1903/4). "Intorno all'ultimo teorema di Fermat". Il Pitagora. 10: 11–13, 41–42. {{cite journal}}: Check date values in: |year= (help)
  73. ^ Werebrusow AS (1905). "On the equation x5 + y5 = Az5 (in Russian)". Moskov. Math. Samml. 25: 466–473.
  74. ^ Rychlik K (1910). "On Fermat's last theorem for n = 5 (in Bohemian)". Časopis Pěst. Mat. 39: 185–195, 305–317.
  75. ^ Terjanian G (1987). "Sur une question de V. A. Lebesgue". Ann. Inst. Fourier. 37: 19–37. doi:10.5802/aif.1096.
  76. ^ Ribenboim, pp. 57–63; Mordell, p. 8; Aczel, p. 44; Singh, p. 106.
  77. ^ Lamé G (1839). "Mémoire sur le dernier théorème de Fermat". C. R. Acad. Sci. Paris. 9: 45–46.
    Lamé G (1840). "Mémoire d'analyse indéterminée démontrant que l'équation x7 + y7 = z7 est impossible en nombres entiers". J. Math. Pures Appl. 5: 195–211.
  78. ^ Lebesgue VA (1840). "Démonstration de l'impossibilité de résoudre l'équation x7 + y7 + z7 = 0 en nombres entiers". J. Math. Pures Appl. 5: 276–279, 348–349.
  79. ^ Freeman L. "Fermat's Last Theorem: Proof for n = 7". Retrieved 23 May 2009.
  80. ^ Genocchi A (1864). "Intorno all'equazioni x7 + y7 + z7 = 0". Annali Mat. 6: 287–288.
    Genocchi A (1874). "Sur l'impossibilité de quelques égalités doubles". C. R. Acad. Sci. Paris. 78: 433–436.
    Genocchi A (1876). "Généralisation du théorème de Lamé sur l'impossibilité de l'équation x7 + y7 + z7 = 0". C. R. Acad. Sci. Paris. 82: 910–913.
  81. ^ Pepin T (1876). "Impossibilité de l'équation x7 + y7 + z7 = 0". C. R. Acad. Sci. Paris. 82: 676–679, 743–747.
  82. ^ Maillet E (1897). "Sur l'équation indéterminée axλt + byλt = czλt". Assoc. Française Avanc. Sci., St. Etienne (sér. II). 26: 156–168.
  83. ^ Thue A (1896). "Über die Auflösbarkeit einiger unbestimmter Gleichungen". Det Kongel. Norske Videnskabers Selskabs Skrifter. 7. Reprinted in Selected Mathematical Papers, pp. 19–30, Oslo:Universitetsforlaget (1977).
  84. ^ Tafelmacher WLA (1897). "La ecuación x3 + y3 = z2: Una demonstración nueva del teorema de fermat para el caso de las sestas potencias". Ann. Univ. Chile, Santiago. 97: 63–80.
  85. ^ Lind B (1909). "Einige zahlentheoretische Sätze". Arch. Math. Phys. 15: 368–369.
  86. ^ a b Kapferer H (1913). "Beweis des Fermatschen Satzes für die Exponenten 6 und 10". Archiv Math. Phys. 21: 143–146.
  87. ^ Swift E (1914). "Solution to Problem 206". Amer. Math. Monthly. 21: 238–239.
  88. ^ a b Breusch R (1960). "A simple proof of Fermat's last theorem for n = 6, n = 10". Math. Mag. 33 (5): 279–281. doi:10.2307/3029800. JSTOR 3029800.
  89. ^ Dirichlet PGL (1832). "Démonstration du théorème de Fermat pour le cas des 14e puissances". J. Reine Angew. Math. 9: 390–393. Reprinted in Werke, vol. I, pp. 189–194, Berlin:G. Reimer (1889); reprinted New York:Chelsea (1969).
  90. ^ Terjanian G (1974). "L'équation x14 + y14 = z14 en nombres entiers". Bull. Sci. Math. (sér. 2). 98: 91–95.
  91. ^ Edwards, pp. 73–74.
  92. ^ a b Edwards, p. 74.
  93. ^ Dickson, p. 733.
  94. ^ Ribenboim P (1979). 13 Lectures on Fermat's Last Theorem. New York: Springer Verlag. pp. 51–54. ISBN 978-0-387-90432-0.
  95. ^ Singh, pp. 97–109.
  96. ^ a b Laubenbacher R, Pengelley D (2007). "Voici ce que j'ai trouvé: Sophie Germain's grand plan to prove Fermat's Last Theorem" (PDF). Retrieved 19 May 2009.
  97. ^ Aczel, p. 57.
  98. ^ Terjanian, G. (1977). "Sur l'équation x2p + y2p = z2p". Comptes rendus hebdomadaires des séances de l'Académie des sciences. Série a et B. 285: 973–975.
  99. ^ Adleman LM, Heath-Brown DR (1985). "The first case of Fermat's last theorem". Inventiones Mathematicae. 79 (2). Berlin: Springer: 409–416. doi:10.1007/BF01388981. {{cite journal}}: Unknown parameter |month= ignored (help)
  100. ^ Aczel, pp. 84–88; Singh, pp. 232–234.
  101. ^ Faltings G (1983). "Endlichkeitssätze für abelsche Varietäten über Zahlkörpern". Inventiones Mathematicae. 73 (3): 349–366. doi:10.1007/BF01388432.
  102. ^ Ribenboim P (1979). 13 Lectures on Fermat's Last Theorem. New York: Springer Verlag. p. 202. ISBN 978-0-387-90432-0.
  103. ^ Wagstaff SS, Jr. (1978). "The irregular primes to 125000". Math. Comp. 32 (142). American Mathematical Society: 583–591. doi:10.2307/2006167. JSTOR 2006167.{{cite journal}}: CS1 maint: multiple names: authors list (link) (PDF)
  104. ^ Buhler J, Crandell R, Ernvall R, Metsänkylä T (1993). "Irregular primes and cyclotomic invariants to four million". Math. Comp. 61 (203). American Mathematical Society: 151–153. doi:10.2307/2152942. JSTOR 2152942.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  105. ^ a b c d e f g h i j k l m n o p q [Fermat's Last Theorem, Simon Singh, 1997, ISBN 1-85702-521-0
  106. ^ Frey G (1986). "Links between stable elliptic curves and certain diophantine equations". Ann. Univ. Sarav. Ser. Math. 1: 1–40.
  107. ^ Singh, pp. 194–198; Aczel, pp. 109–114.
  108. ^ Singh, p. 205; Aczel, pp. 117–118.
  109. ^ Singh, pp. 237–238; Aczel, pp. 121–122.
  110. ^ Singh, pp. 239–243; Aczel, pp. 122–125.
  111. ^ Singh, pp. 244–253; Aczel, pp. 1–4, 126–128.
  112. ^ Aczel, pp. 128–130.
  113. ^ Singh, p. 257.
  114. ^ Singh, pp. 269–274.
  115. ^ Singh, pp. 275–277; Aczel, pp. 132–134.
  116. ^ Wiles, Andrew (1995). "Modular elliptic curves and Fermat's Last Theorem" (PDF). Annals of Mathematics. 141 (3). Annals of Mathematics: 443–551. doi:10.2307/2118559. JSTOR 2118559. OCLC 37032255.
  117. ^ Taylor R, Wiles A (1995). "Ring theoretic properties of certain Hecke algebras". Annals of Mathematics. 141 (3). Annals of Mathematics: 553–572. doi:10.2307/2118560. JSTOR 2118560. OCLC 37032255.
  118. ^ Lenstra, Jr. H.W. (1992). On the inverse Fermat equation. Discrete Mathematics, 106–107, pp. 329–331.
  119. ^ Newton, M., "A radical diophantine equation", Journal of Number Theory 13 (1981), 495–498.
  120. ^ Bennett, Curt D., Glass, Andrew M.W., and Székely, Gábor J. (2004). Fermat’s last theorem for rational exponents. The American Mathematical Monthly, 111, no. 4, pp. 322–329.
  121. ^ Dickson, pp. 688–691
  122. ^ Voles, Roger, "Integer solutions of ," Mathematical Gazette 83, July 1999, 269–271.
  123. ^ Richinick, Jennifer, "The upside-down Pythagorean Theorem," Mathematical Gazette 92, July 2008, 313–317.
  124. ^ BBC Documentary
  125. ^ Avigad, Jeremy (2003). "Number theory and elementary arithmetic". Philosophia Mathematica. Philosophy of Mathematics, its Learning, and its Application. Series III. 11 (3): 257–284. doi:10.1093/philmat/11.3.257. ISSN 0031-8019. MR 2006194. CiteSeerx10.1.1.105.6509.
  126. ^ Aczel, p. 69; Singh, p. 105.
  127. ^ Aczel, p. 69.
  128. ^ a b Koshy T (2001). Elementary number theory with applications. New York: Academic Press. p. 544. ISBN 978-0-12-421171-1.
  129. ^ Singh, pp. 120–125, 131–133, 295–296; Aczel, p. 70.
  130. ^ Singh, pp. 120–125.
  131. ^ Singh, p. 284
  132. ^ Singh, p. 295.
  133. ^ Singh, pp. 295–296.

Bibliography

Further reading

External links