Fine Art

.

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 in the margin of a copy of Arithmetica where he claimed he had a proof that was too large to fit in the margin. The first successful proof was released in 1994 by Andrew Wiles, and formally published in 1995, after 358 years of effort by mathematicians. 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 notable theorems in the history of mathematics and prior to its proof it was in the Guinness Book of World Records for "most difficult mathematical problems".

1 Overview
1.1 Subsequent developments and solution
2 Mathematical history
2.1 Pythagoras and Diophantus
2.1.1 Pythagorean triples
2.1.2 Diophantine equations
2.2 Fermat's conjecture
2.3 Proofs for specific exponents
2.3.1 Sophie Germain
2.3.2 Ernst Kummer and the theory of ideals
2.3.3 Mordell conjecture
2.3.4 Computational studies
2.4 Connection with elliptic curves
2.4.1 Taniyama–Shimura–Weil conjecture
2.4.2 Frey's equation / Ribet's theorem
2.4.3 Wiles's general proof
2.5 Subsequent developments
3 Exponents other than positive integers
3.1 Reciprocal Integers (Inverse Fermat Equation)
3.2 Rational exponents
3.3 Negative exponents
3.3.1 n = –1
3.3.2 n = –2
3.3.3 Integer n < –2
4 Base values other than positive integers
5 Monetary prizes
6 See also
7 Footnotes
8 References
9 Bibliography
10 Further reading

Overview

Fermat's Last Theorem stood as an unsolved riddle in mathematics for over three and a half centuries. The theorem itself is a deceptively simple statement that Fermat stated he had proved around 1637. His claim was discovered some 30 years later, after his death, written in the margin of a book, but with no proof provided.

The claim eventually became one of the most notable unsolved problems of mathematics. Attempts to prove it prompted substantial development in number theory, and over time Fermat's Last Theorem gained prominence as an unsolved problem in mathematics. It is based on the Pythagorean theorem, which states that a2 + b2 = c2, where a and b are the lengths of the legs of a right triangle and c is the length of the hypotenuse.

The Pythagorean equation has an infinite number of positive integer solutions for a, b, and c; these solutions are known as Pythagorean triples. Fermat stated that the more general equation an + bn = cn had no solutions in positive integers, if n is an integer greater than 2. 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 that 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 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 357 years later by Andrew Wiles in 1994, 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 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's work, between 1996 and 2001.


Mathematical history
Pythagoras and Diophantus
Pythagorean triples
Main article: Pythagorean triple

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]

\( a^2 + b^2 = c^2.\ \)

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 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:

\( A = x + y\ \)
\( B = x^2 + y^2.\ \)

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] that 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 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 marvellous proof of this, which this margin is too narrow to contain.[14][15]

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.[16] The margin note became known as Fermat’s Last Theorem,[17] as it was the last of Fermat’s asserted theorems to remain unproven.[18]

It is not known whether Fermat had actually found a valid proof for all exponents n, but it appears unlikely. Only one related proof by him has survived, namely for the case n = 4, as described in the section Proofs for specific exponents. While 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,[19] he never posed the general case.[20] Moreover, in the last thirty years of his life, Fermat never again wrote of his “truly marvellous proof” of the general case, and never published it. Van der Poorten[21] suggests that while the absence of a proof is insignificant, the lack of challenges means Fermat realised he did not have a proof; he quotes Weil[22] as saying Fermat must have briefly deluded himself with an irretrievable idea.

The techniques Fermat might have used in such a “marvellous proof” are unknown.

Taylor and Wiles’s proof relies on 20th century techniques.[23] Fermat’s proof would have had to have been elementary by comparison, given the mathematical knowledge of his time.

While Harvey Friedman’s grand conjecture implies that any provable theorem (including Fermat’s last theorem) can be proved using only ‘elementary function arithmetic’, such a proof need only be ‘elementary’ in a technical sense but could involve millions of steps, and thus be far too long to have been Fermat’s proof.


Proofs for specific exponents
Main article: Proof of Fermat's Last Theorem for specific exponents
Fermat's infinite descent for Fermat's Last Theorem case n=4 in the 1670 edition of the Arithmetica of Diophantus (p. p. 338-339).

Only one relevant proof by Fermat has survived, in which he 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.[24][25] His proof is equivalent to demonstrating that the equation

\( x^4 - y^4 = z^2 \)

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 equationa4 + b4 = c4 can be written as c4b4 = (a2)2.

Alternative proofs of the case n = 4 were developed later[26] by Frénicle de Bessy (1676),[27] Leonhard Euler (1738),[28] Kausler (1802),[29] Peter Barlow (1811),[30] Adrien-Marie Legendre (1830),[31] Schopis (1825),[32] Terquem (1846),[33] Joseph Bertrand (1851),[34] Victor Lebesgue (1853, 1859, 1862),[35] Theophile Pepin (1883),[36] Tafelmacher (1893),[37] David Hilbert (1897),[38] Bendz (1901),[39] Gambioli (1901),[40] Leopold Kronecker (1901),[41] Bang (1905),[42] Sommer (1907),[43] Bottari (1908),[44] Karel Rychlík (1910),[45] Nutzhorn (1912),[46] Robert Carmichael (1913),[47] Hancock (1931),[48] and Vrǎnceanu (1966).[49]All proofs for specific exponents used Fermat's technique of infinite descent,[citation needed] either in its original form, or in the form of descent on elliptic curves or abelian varieties. The details and auxiliary arguments, however, were often ad hoc and tied to the individual exponent under consideration.[97] Since they became ever more complicated as p increased, it seemed unlikely that the general case of Fermat's Last Theorem could be proved by building upon the proofs for individual exponents.[97] Although some general results on Fermat's Last Theorem were published in the early 19th century by Niels Henrik Abel and Peter Barlow,[98][99] the first significant work on the general theorem was done by Sophie Germain.[100]
Sophie Germain

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),[50] Barbara (2007),[51] and Dolan (2011).[52]

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.[53] 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 would suffice to prove that it has no solutions for at least one prime factor of every n. Each integer n > 2 is divisible by 4 or an odd prime number (or both). Therefore, Fermat's Last Theorem could be proved for all n if it could be proved for n = 4 and for all odd primes p.

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.[54] In 1770, Leonhard Euler gave a proof of p = 3,[55] but his proof by infinite descent[56] contained a major gap.[57] However, since Euler himself had proven the lemma necessary to complete the proof in other work, he is generally credited with the first proof.[58] Independent proofs were published[59] by Kausler (1802),[29] Legendre (1823, 1830),[31][60] Calzolari (1855),[61] Gabriel Lamé (1865),[62] Peter Guthrie Tait (1872),[63] Günther (1878),[64] Gambioli (1901),[40] Krey (1909),[65] Rychlík (1910),[45] Stockhaus (1910),[66] Carmichael (1915),[67] Johannes van der Corput (1915),[68] Axel Thue (1917),[69] and Duarte (1944).[70] The case p = 5 was proven[71] independently by Legendre and Peter Gustav Lejeune Dirichlet around 1825.[72] Alternative proofs were developed[73] by Carl Friedrich Gauss (1875, posthumous),[74] Lebesgue (1843),[75] Lamé (1847),[76] Gambioli (1901),[40][77] Werebrusow (1905),[78] Rychlík (1910),[79] van der Corput (1915),[68] and Guy Terjanian (1987).[80] The case p = 7 was proven[81] by Lamé in 1839.[82] His rather complicated proof was simplified in 1840 by Lebesgue,[83] and still simpler proofs[84] were published by Angelo Genocchi in 1864, 1874 and 1876.[85] Alternative proofs were developed by Théophile Pépin (1876)[86] and Edmond Maillet (1897).[87]

Fermat's Last Theorem was also proven for the exponents n = 6, 10, and 14. Proofs for n = 6 were published by Kausler,[29] Thue,[88] Tafelmacher,[89] Lind,[90] Kapferer,[91] Swift,[92] and Breusch.[93] Similarly, Dirichlet[94] and Terjanian[95] each proved the case n = 14, while Kapferer[91] and Breusch[93] 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.[96]

All proofs for specific exponents used Fermat's technique of infinite descent,[citation needed] either in its original form, or in the form of descent on elliptic curves or abelian varieties. The details and auxiliary arguments, however, were often ad hoc and tied to the individual exponent under consideration.[97] Since they became ever more complicated as p increased, it seemed unlikely that the general case of Fermat's Last Theorem could be proved by building upon the proofs for individual exponents.[97] Although some general results on Fermat's Last Theorem were published in the early 19th century by Niels Henrik Abel and Peter Barlow,[98][99] the first significant work on the general theorem was done by Sophie Germain.[100]


Main article: Sophie Germain

In the early 19th century, Sophie Germain developed several novel approaches to prove Fermat's Last Theorem for all exponents.[101] 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.[101][102] 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.[103] 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.[104]

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) that 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.[105] This conjecture was proven in 1983 by Gerd Faltings,[106] 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.[107] By 1978, Samuel Wagstaff had extended this to all primes less than 125,000.[108] By 1993, Fermat's Last Theorem had been proven for all primes less than four million.[109]

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"[110]:211 Taniyama–Shimura-Weil conjecture, proposed around 1955—which many mathematicians believed would be near to impossible to prove,[110]:223 and 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 1994, 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.
Taniyama–Shimura–Weil conjecture


Main article: Modularity theorem

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 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.[110]:211–215

Even after gaining serious attention, the conjecture was seen by contemporary mathematicians as extraordinarily difficult or perhaps inaccessible to proof.[110]:203–205, 223, 226 For example, Wiles' ex-supervisor John Coates states that it seemed "impossible to actually prove",[110]: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]."[110]:223
Frey's equation / Ribet's theorem


Main articles: Frey curve and 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.[111] 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[112] and 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) that included Frey's equation – and this was widely believed inaccessible to proof by contemporary mathematicians.[110]: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
Main articles: Andrew Wiles and Wiles's proof of Fermat's Last Theorem

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.[113]

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.[110]:229–230 His initial study suggested proof by induction,[110]:230–232, 249–252 and he based his initial work and first significant breakthrough on Galois theory[110]: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.[110]:258–259 However, by the summer of 1991, Iwasawa theory also seemed to not be reaching the central issues in the problem.[110]:259–260[114] 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 that seemed "tailor made" for the inductive part of his proof.[110]: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 Wiles used seemed to work correctly.[110]:261–265[115]

By mid-May 1993 Wiles felt able to tell his wife he thought he had solved the proof of Fermat's Last Theorem,[110]: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.[116] 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),[117] who alerted Wiles on 23 August 1993.[118]

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.[110]: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.[119]

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.[120] On 24 October 1994, Wiles submitted two manuscripts, "Modular elliptic curves and Fermat's Last Theorem"[121] and "Ring theoretic properties of certain Hecke algebras",[122] the second of which was co-authored with Taylor and proved that certain conditions were met that 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 proven.[123][124][125] 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
Reciprocal Integers (Inverse Fermat Equation)

The equation \( a^{1/m} + b^{1/m} = c^{1/m} \) can be considered the "inverse" Fermat equation. All solutions of this equation were computed by Lenstra in 1992.[126] In the case in which the mth roots are required to be real and positive, all solutions are given by[127]

\( a=rs^m \)
\( b=rt^m \)
\( c=r(s+t)^m \)

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

For the Diophantine equation \( a^{n/m} + b^{n/m} = c^{n/m} \) with n not equal to 1, Bennett, Glass, and Székely proved in 2004 for n > 2 that if n and m are coprime, then there are integer solutions if and only if 6 divides m, and \( a^{1/m}, b^{1/m} \), and \( c^{1/m} \) are different complex 6th roots of the same real number.[128]
Negative exponents
n = –1

All primitive (pairwise coprime) integer solutions to the optic equation a^{-1}+b^{-1}=c^{-1} can be written as[129]

\( a=mn+m^2, \)
\( b=mn+n^2, \)
\( c=mn \)

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.[130][131] All primitive solutions to \( a^{-2}+b^{-2}=d^{-2} \) are given by

\( a=(v^2-u^2)(v^2+u^2), \, \)
\( b=2uv(v^2+u^2), \, \)
\( d=2uv(v^2-u^2), \, \)

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

\( c=(v^2+u^2)^2, \, \)

so (a, b, c) is a Pythagorean triple.
Integer n < –2

There are no solutions in integers for \( ^n+b^n=c^n \) for integers n < –2. If there were, the equation could be multiplied through by \( a^{|n|}b^{|n|}c^{|n|} \) to obtain \( (bc)^{|n|}+(ac)^{|n|}=(ab)^{|n|} \), which is impossible by Fermat's Last Theorem.
Base values other than positive integers

Fermat's last theorem can easily be extended to positive rationals:

\( \left(\frac{a}{x}\right)^n+\left(\frac{b}{y}\right)^n=\left(\frac{c}{z}\right)^n \)

can have no solutions with n > 2, because any solution could be rearranged as:

\( (ayz)^n+(bxz)^n=(cxy)^n, \)

to which Fermat's Last Theorem applies.
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.[132] 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.[133] Another prize was offered in 1883 by the Academy of Brussels.[134]

In 1908, the German industrialist and amateur mathematician Paul Wolfskehl bequeathed 100,000 gold marks—a large sum at the time—to the Göttingen Academy of Sciences to offer as a prize for a complete proof of Fermat's Last Theorem.[135] 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.[136] Wiles collected the Wolfskehl prize money, then worth $50,000, on 27 June 1997.[137]

Prior to Wiles's proof, thousands of incorrect proofs were submitted to the Wolfskehl committee, amounting to roughly 10 feet (3 meters) of correspondence.[138] 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".[139] 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."[134]
See also

Beal's conjecture
Diophantus II.VIII
Euler's sum of powers conjecture
Fermat's Last Theorem in fiction
Proof of impossibility
Sophie Germain prime
Sums of powers, a list of related conjectures and theorems
Wall–Sun–Sun prime

Footnotes

If the exponent "n" were not prime or 4, then it would be possible to write n either as a product of two smaller integers (n = P*Q) in which P is a prime number greater than 2, 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 that is smaller than N, as well; or else as n would be a power of 2 greater than four and writing n=4*Q, the same argument would hold.
For example, \( \left((j^r+1)^s\right)^r + \left(j(j^r+1)^s)\right)^r = (j^r+1)^{rs+1}. \)

This elliptic curve was first suggested in the 1960s by Yves Hellegouarch (de), 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

Stark, pp. 151–155.
Stillwell J (2003). Elements of Number Theory. New York: Springer-Verlag. pp. 110–112. ISBN 0-387-95587-9.
Aczel, pp. 13–15
Singh, pp. 18–20.
Singh, p. 6.
Stark, pp. 145–146.
Singh, pp. 50–51.
Stark, p. 145.
Aczel, pp. 44–45; Singh, pp. 56–58.
Aczel, pp. 14–15.
Stark, pp. 44–47.
Friberg, pp. 333– 334.
Dickson, p. 731; Singh, pp. 60–62; Aczel, p. 9.
T. Heath, Diophantus of Alexandria Second Edition, Cambridge University Press, 1910, reprinted by Dover, NY, 1964, pp. 144-145
Panchishkin, p. 341
Singh, pp. 62–66.
Dickson, p. 731.
Singh, p. 67; Aczel, p. 10.
Ribenboim, pp. 13, 24.
van der Poorten, Notes and Remarks 1.2, p. 5.
van der Poorten, loc. cit.
André Weil (1984). Number Theory: An approach through history. From Hammurapi to Legendre. Basel, Switzerland: Birkhäuser. p. 104.
BBC Documentary
Freeman L. "Fermat's One Proof". Retrieved 23 May 2009.
Dickson, pp. 615–616; Aczel, p. 44.
Ribenboim, pp. 15–24.
Frénicle de Bessy, Traité des Triangles Rectangles en Nombres, vol. I, 1676, Paris. Reprinted in Mém. Acad. Roy. Sci., 5, 1666–1699 (1729).
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).
Kausler CF (1802). "Nova demonstratio theorematis nec summam, nec differentiam duorum cuborum cubum esse posse". Novi Acta Acad. Petrop. 13: 245–253.
Barlow P (1811). An Elementary Investigation of Theory of Numbers. St. Paul's Church-Yard, London: J. Johnson. pp. 144–145.
Legendre AM (1830). Théorie des Nombres (Volume II) (3rd ed.). Paris: Firmin Didot Frères. Reprinted in 1955 by A. Blanchard (Paris).
Schopis (1825). Einige Sätze aus der unbestimmten Analytik. Gummbinnen: Programm.
Terquem O (1846). "Théorèmes sur les puissances des nombres". Nouv. Ann. Math. 5: 70–87.
Bertrand J (1851). Traité Élémentaire d'Algèbre. Paris: Hachette. pp. 217–230, 395.
Lebesgue VA (1853). "Résolution des équations biquadratiques z2 = x4 ± 2my4, z2 = 2mx4 − y4, 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.
Pepin T (1883). "Étude sur l'équation indéterminée ax4 + by4 = cz2". Atti Accad. Naz. Lincei 36: 34–70.
Tafelmacher WLA (1893). "Sobre la ecuación x4 + y4 = z4". Ann. Univ. Chile 84: 307–320.
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.
Bendz TR (1901). Öfver diophantiska ekvationen xn + yn = zn. Uppsala: Almqvist & Wiksells Boktrycken.
Gambioli D (1901). "Memoria bibliographica sull'ultimo teorema di Fermat". Period. Mat. 16: 145–192.
Kronecker L (1901). Vorlesungen über Zahlentheorie, vol. I. Leipzig: Teubner. pp. 35–38. Reprinted by New York:Springer-Verlag in 1978.
Bang A (1905). "Nyt Bevis for at Ligningen x4 − y4 = z4, ikke kan have rationale Løsinger". Nyt Tidsskrift Mat. 16B: 35–36.
Sommer J (1907). Vorlesungen über Zahlentheorie. Leipzig: Teubner.
Bottari A (1908). "Soluzione intere dell'equazione pitagorica e applicazione alla dimostrazione di alcune teoremi della teoria dei numeri". Period. Mat. 23: 104–110.
Rychlik K (1910). "On Fermat's last theorem for n = 4 and n = 3 (in Bohemian)". Časopis Pěst. Mat. 39: 65–86.
Nutzhorn F (1912). "Den ubestemte Ligning x4 + y4 = z4". Nyt Tidsskrift Mat. 23B: 33–38.
Carmichael RD (1913). "On the impossibility of certain Diophantine equations and systems of equations". Amer. Math. Monthly (Mathematical Association of America) 20 (7): 213–221. doi:10.2307/2974106. JSTOR 2974106.
Hancock H (1931). Foundations of the Theory of Algebraic Numbers, vol. I. New York: Macmillan.
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.
Grant, Mike, and Perella, Malcolm, "Descending to the irrational", Mathematical Gazette 83, July 1999, pp.263–267.
Barbara, Roy, "Fermat's last theorem in the case n=4", Mathematical Gazette 91, July 2007, 260–262.
Dolan, Stan, "Fermat's method of descente infinie", Mathematical Gazette 95, July 2011, 269–271.
Ribenboim, pp. 1–2.
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.
Euler L (1770) Vollständige Anleitung zur Algebra, Roy. Acad. Sci., St. Petersburg.
Freeman L. "Fermat's Last Theorem: Proof for n = 3". Retrieved 23 May 2009.
Ribenboim, pp. 24–25; Mordell, pp. 6–8; Edwards, pp. 39–40.
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.
Ribenboim, pp. 33, 37–41.
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.
Calzolari L (1855). Tentativo per dimostrare il teorema di Fermat sull'equazione indeterminata xn + yn = zn. Ferrara.
Lamé G (1865). "Étude des binômes cubiques x3 ± y3". C. R. Acad. Sci. Paris 61: 921–924, 961–965.
Tait PG (1872). "Mathematical Notes". Proc. Roy. Soc. Edinburgh 7: 144.
Günther S (1878). "Über die unbestimmte Gleichung x3 + y3 = z3". Sitzungsberichte Böhm. Ges. Wiss.: 112–120.
Krey H (1909). "Neuer Beweis eines arithmetischen Satzes". Math. Naturwiss. Blätter 6: 179–180.
Stockhaus H (1910). Beitrag zum Beweis des Fermatschen Satzes. Leipzig: Brandstetter.
Carmichael RD (1915). Diophantine Analysis. New York: Wiley.
van der Corput JG (1915). "Quelques formes quadratiques et quelques équations indéterminées". Nieuw Archief Wisk. 11: 45–75.
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.
Duarte FJ (1944). "Sobre la ecuación x3 + y3 + z3 = 0". Ciencias Fis. Mat. Naturales (Caracas) 8: 971–979.
Freeman L. "Fermat's Last Theorem: Proof for n = 5". Retrieved 23 May 2009.
Ribenboim, p. 49; Mordell, p. 8–9; Aczel, p. 44; Singh, p. 106.
Ribenboim, pp. 55–57.
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. Check date values in: |date= (help)
Lebesgue VA (1843). "Théorèmes nouveaux sur l'équation indéterminée x5 + y5 = az5". J. Math. Pures Appl. 8: 49–70.
Lamé G (1847). "Mémoire sur la résolution en nombres complexes de l'équation A5 + B5 + C5 = 0". J. Math. Pures Appl. 12: 137–171.
Gambioli D (1903/4). "Intorno all'ultimo teorema di Fermat". Il Pitagora 10: 11–13, 41–42. Check date values in: |date= (help)
Werebrusow AS (1905). "On the equation x5 + y5 = Az5 (in Russian)". Moskov. Math. Samml. 25: 466–473.
Rychlik K (1910). "On Fermat's last theorem for n = 5 (in Bohemian)". Časopis Pěst. Mat. 39: 185–195, 305–317.
Terjanian G (1987). "Sur une question de V. A. Lebesgue". Ann. Inst. Fourier 37: 19–37. doi:10.5802/aif.1096.
Ribenboim, pp. 57–63; Mordell, p. 8; Aczel, p. 44; Singh, p. 106.
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.
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.
Freeman L. "Fermat's Last Theorem: Proof for n = 7". Retrieved 23 May 2009.
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.
Pepin T (1876). "Impossibilité de l'équation x7 + y7 + z7 = 0". C. R. Acad. Sci. Paris 82: 676–679, 743–747.
Maillet E (1897). "Sur l'équation indéterminée axλt + byλt = czλt". Assoc. Française Avanc. Sci., St. Etienne (sér. II) 26: 156–168.
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).
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.
Lind B (1909). "Einige zahlentheoretische Sätze". Arch. Math. Phys. 15: 368–369.
Kapferer H (1913). "Beweis des Fermatschen Satzes für die Exponenten 6 und 10". Archiv Math. Phys. 21: 143–146.
Swift E (1914). "Solution to Problem 206". Amer. Math. Monthly 21: 238–239.
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.
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).
Terjanian G (1974). "L'équation x14 + y14 = z14 en nombres entiers". Bull. Sci. Math. (sér. 2) 98: 91–95.
Edwards, pp. 73–74.
Edwards, p. 74.
Dickson, p. 733.
Ribenboim P (1979). 13 Lectures on Fermat's Last Theorem. New York: Springer Verlag. pp. 51–54. ISBN 978-0-387-90432-0.
Singh, pp. 97–109.
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.
Aczel, p. 57.
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.
Adleman LM, Heath-Brown DR (June 1985). "The first case of Fermat's last theorem". Inventiones Mathematicae (Berlin: Springer) 79 (2): 409–416. doi:10.1007/BF01388981.
Aczel, pp. 84–88; Singh, pp. 232–234.
Faltings G (1983). "Endlichkeitssätze für abelsche Varietäten über Zahlkörpern". Inventiones Mathematicae 73 (3): 349–366. doi:10.1007/BF01388432.
Ribenboim P (1979). 13 Lectures on Fermat's Last Theorem. New York: Springer Verlag. p. 202. ISBN 978-0-387-90432-0.
Wagstaff SS, Jr. (1978). "The irregular primes to 125000". Math. Comp. (American Mathematical Society) 32 (142): 583–591. doi:10.2307/2006167. JSTOR 2006167. (PDF)
Buhler J, Crandell R, Ernvall R, Metsänkylä T (1993). "Irregular primes and cyclotomic invariants to four million". Math. Comp. (American Mathematical Society) 61 (203): 151–153. doi:10.2307/2152942. JSTOR 2152942.
[Fermat's Last Theorem, Simon Singh, 1997, ISBN 1-85702-521-0
Frey G (1986). "Links between stable elliptic curves and certain diophantine equations". Ann. Univ. Sarav. Ser. Math. 1: 1–40.
Singh, pp. 194–198; Aczel, pp. 109–114.
Singh, p. 205; Aczel, pp. 117–118.
Singh, pp. 237–238; Aczel, pp. 121–122.
Singh, pp. 239–243; Aczel, pp. 122–125.
Singh, pp. 244–253; Aczel, pp. 1–4, 126–128.
Aczel, pp. 128–130.
Singh, p. 257.
Singh, pp. 269–274.
Singh, pp. 275–277; Aczel, pp. 132–134.
Wiles, Andrew (1995). "Modular elliptic curves and Fermat's Last Theorem" (PDF). Annals of Mathematics 141 (3): 443–551. doi:10.2307/2118559. JSTOR 2118559. OCLC 37032255.
Taylor R, Wiles A (1995). "Ring theoretic properties of certain Hecke algebras". Annals of Mathematics 141 (3): 553–572. doi:10.2307/2118560. JSTOR 2118560. OCLC 37032255.
Diamond, Fred (1996). "On deformation rings and Hecke rings". Annals of Mathematics. Second Series 144 (1): 137–166. doi:10.2307/2118586. ISSN 0003-486X. MR 1405946.
Conrad, Brian; Diamond, Fred; Taylor, Richard (1999). "Modularity of certain potentially Barsotti-Tate Galois representations". Journal of the American Mathematical Society 12 (2): 521–567. doi:10.1090/S0894-0347-99-00287-8. ISSN 0894-0347. MR 1639612.
Breuil, Christophe; Conrad, Brian; Diamond, Fred; Taylor, Richard (2001). "On the modularity of elliptic curves over Q: wild 3-adic exercises". Journal of the American Mathematical Society 14 (4): 843–939. doi:10.1090/S0894-0347-01-00370-8. ISSN 0894-0347. MR 1839918.
Lenstra, Jr. H.W. (1992). On the inverse Fermat equation, Discrete Mathematics, 106–107, pp. 329–331.
Newton, M., "A radical diophantine equation", Journal of Number Theory 13 (1981), 495–498.
Bennett, Curtis D.; Glass, A. M. W.; Székely, Gábor J. (2004). "Fermat's last theorem for rational exponents". American Mathematical Monthly 111 (4): 322–329. doi:10.2307/4145241. MR 2057186.
Dickson, pp. 688–691
Voles, Roger (July 1999). "Integer solutions of a−2 + b−2 = d−2". Mathematical Gazette 83: 269–271.
Richinick, Jennifer (July 2008). "The upside-down Pythagorean Theorem". Mathematical Gazette 92: 313–317.
Aczel, p. 69; Singh, p. 105.
Aczel, p. 69.
Koshy T (2001). Elementary number theory with applications. New York: Academic Press. p. 544. ISBN 978-0-12-421171-1.
Singh, pp. 120–125, 131–133, 295–296; Aczel, p. 70.
Singh, pp. 120–125.
Singh, p. 284
Singh, p. 295.

Singh, pp. 295–296.

Bibliography

Aczel, Amir (30 September 1996). Fermat's Last Theorem: Unlocking the Secret of an Ancient Mathematical Problem. Four Walls Eight Windows. ISBN 978-1-56858-077-7.
Dickson LE (1919). History of the Theory of Numbers. Volume II. Diophantine Analysis. New York: Chelsea Publishing. pp. 545–550, 615–621, 688–691, 731–776.
Edwards, HM (1997). Fermat's Last Theorem. A Genetic Introduction to Algebraic Number Theory. Graduate Texts in Mathematics 50. New York: Springer-Verlag.
Friberg, Joran (2007). Amazing Traces of a Babylonian Origin in Greek Mathematics. World Scientific Publishing Company. ISBN 978-981-270-452-8.
Kleiner I (2000). "From Fermat to Wiles: Fermat's Last Theorem Becomes a Theorem" (PDF). Elem. Math. 55: 19–37. doi:10.1007/PL00000079.
Mordell LJ (1921). Three Lectures on Fermat's Last Theorem. Cambridge: Cambridge University Press.
Panchishkin, Alekseĭ Alekseevich (2007). Introduction to Modern Number Theory (Encyclopedia of Mathematical Sciences. Springer Berlin Heidelberg New York. ISBN 978-3-540-20364-3.
Ribenboim P (2000). Fermat's Last Theorem for Amateurs. New York: Springer-Verlag. ISBN 978-0-387-98508-4.
Singh S (October 1998). Fermat's Enigma. New York: Anchor Books. ISBN 978-0-385-49362-8.
Stark H (1978). An Introduction to Number Theory. MIT Press. ISBN 0-262-69060-8.

Further reading

Bell, Eric T. (6 August 1998) [1961]. The Last Problem. New York: The Mathematical Association of America. ISBN 978-0-88385-451-8.
Benson, Donald C. (5 April 2001). The Moment of Proof: Mathematical Epiphanies. Oxford University Press. ISBN 978-0-19-513919-8.
Brudner, Harvey J. (1994). Fermat and the Missing Numbers. WLC, Inc. ISBN 978-0-9644785-0-3.
Edwards, H. M. (March 1996) [1977]. Fermat's Last Theorem. New York: Springer-Verlag. ISBN 978-0-387-90230-2.
Faltings G (July 1995). "The Proof of Fermat's Last Theorem by R. Taylor and A. Wiles" (PDF). Notices of the AMS 42 (7): 743–746. ISSN 0002-9920.
Mozzochi, Charles (7 December 2000). The Fermat Diary. American Mathematical Society. ISBN 978-0-8218-2670-6.
Ribenboim P (1979). 13 Lectures on Fermat's Last Theorem. New York: Springer Verlag. ISBN 978-0-387-90432-0.
van der Poorten, Alf (6 March 1996). Notes on Fermat's Last Theorem. WileyBlackwell. ISBN 978-0-471-06261-5.
Saikia, Manjil P (July 2011). "A Study of Kummer's Proof of Fermat's Last Theorem for Regular Primes" (PDF). IISER Mohali (India) Summer Project Report.

External links
Overview

Fermat's Last Theorem stood as an unsolved riddle in mathematics for over three and a half centuries. The theorem itself is a deceptively simple statement that Fermat stated he had proved around 1637. His claim was discovered some 30 years later, after his death, written in the margin of a book, but with no proof provided.

The claim eventually became one of the most notable unsolved problems of mathematics. Attempts to prove it prompted substantial development in number theory, and over time Fermat's Last Theorem gained prominence as an unsolved problem in mathematics. It is based on the Pythagorean theorem, which states that a2 + b2 = c2, where a and b are the lengths of the legs of a right triangle and c is the length of the hypotenuse.

The Pythagorean equation has an infinite number of positive integer solutions for a, b, and c; these solutions are known as Pythagorean triples. Fermat stated that the more general equation an + bn = cn had no solutions in positive integers, if n is an integer greater than 2. 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 that 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 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 357 years later by Andrew Wiles in 1994, 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 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's work, between 1996 and 2001.
Mathematical history
Pythagoras and Diophantus
Pythagorean triples
Main article: Pythagorean triple

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]

a^2 + b^2 = c^2.\

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 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:

A = x + y\
B = x^2 + y^2.\

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] that 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 margin that 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 marvellous proof of this, which this margin is too narrow to contain.[14][15]

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.[16] The margin note became known as Fermat’s Last Theorem,[17] as it was the last of Fermat’s asserted theorems to remain unproven.[18]

It is not known whether Fermat had actually found a valid proof for all exponents n, but it appears unlikely. Only one related proof by him has survived, namely for the case n = 4, as described in the section Proofs for specific exponents. While 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,[19] he never posed the general case.[20] Moreover, in the last thirty years of his life, Fermat never again wrote of his “truly marvellous proof” of the general case, and never published it. Van der Poorten[21] suggests that while the absence of a proof is insignificant, the lack of challenges means Fermat realised he did not have a proof; he quotes Weil[22] as saying Fermat must have briefly deluded himself with an irretrievable idea.

The techniques Fermat might have used in such a “marvellous proof” are unknown.

Taylor and Wiles’s proof relies on 20th century techniques.[23] Fermat’s proof would have had to have been elementary by comparison, given the mathematical knowledge of his time.

While Harvey Friedman’s grand conjecture implies that any provable theorem (including Fermat’s last theorem) can be proved using only ‘elementary function arithmetic’, such a proof need only be ‘elementary’ in a technical sense but could involve millions of steps, and thus be far too long to have been Fermat’s proof.
Proofs for specific exponents
Main article: Proof of Fermat's Last Theorem for specific exponents
Fermat's infinite descent for Fermat's Last Theorem case n=4 in the 1670 edition of the Arithmetica of Diophantus (p. p. 338-339).

Only one relevant proof by Fermat has survived, in which he 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.[24][25] His proof is equivalent to demonstrating that the equation

x^4 - y^4 = z^2

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[26] by Frénicle de Bessy (1676),[27] Leonhard Euler (1738),[28] Kausler (1802),[29] Peter Barlow (1811),[30] Adrien-Marie Legendre (1830),[31] Schopis (1825),[32] Terquem (1846),[33] Joseph Bertrand (1851),[34] Victor Lebesgue (1853, 1859, 1862),[35] Theophile Pepin (1883),[36] Tafelmacher (1893),[37] David Hilbert (1897),[38] Bendz (1901),[39] Gambioli (1901),[40] Leopold Kronecker (1901),[41] Bang (1905),[42] Sommer (1907),[43] Bottari (1908),[44] Karel Rychlík (1910),[45] Nutzhorn (1912),[46] Robert Carmichael (1913),[47] Hancock (1931),[48] and Vrǎnceanu (1966).[49]

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),[50] Barbara (2007),[51] and Dolan (2011).[52]

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.[53] 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 would suffice to prove that it has no solutions for at least one prime factor of every n. Each integer n > 2 is divisible by 4 or an odd prime number (or both). Therefore, Fermat's Last Theorem could be proved for all n if it could be proved for n = 4 and for all odd primes p.

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.[54] In 1770, Leonhard Euler gave a proof of p = 3,[55] but his proof by infinite descent[56] contained a major gap.[57] However, since Euler himself had proven the lemma necessary to complete the proof in other work, he is generally credited with the first proof.[58] Independent proofs were published[59] by Kausler (1802),[29] Legendre (1823, 1830),[31][60] Calzolari (1855),[61] Gabriel Lamé (1865),[62] Peter Guthrie Tait (1872),[63] Günther (1878),[64] Gambioli (1901),[40] Krey (1909),[65] Rychlík (1910),[45] Stockhaus (1910),[66] Carmichael (1915),[67] Johannes van der Corput (1915),[68] Axel Thue (1917),[69] and Duarte (1944).[70] The case p = 5 was proven[71] independently by Legendre and Peter Gustav Lejeune Dirichlet around 1825.[72] Alternative proofs were developed[73] by Carl Friedrich Gauss (1875, posthumous),[74] Lebesgue (1843),[75] Lamé (1847),[76] Gambioli (1901),[40][77] Werebrusow (1905),[78] Rychlík (1910),[79] van der Corput (1915),[68] and Guy Terjanian (1987).[80] The case p = 7 was proven[81] by Lamé in 1839.[82] His rather complicated proof was simplified in 1840 by Lebesgue,[83] and still simpler proofs[84] were published by Angelo Genocchi in 1864, 1874 and 1876.[85] Alternative proofs were developed by Théophile Pépin (1876)[86] and Edmond Maillet (1897).[87]

Fermat's Last Theorem was also proven for the exponents n = 6, 10, and 14. Proofs for n = 6 were published by Kausler,[29] Thue,[88] Tafelmacher,[89] Lind,[90] Kapferer,[91] Swift,[92] and Breusch.[93] Similarly, Dirichlet[94] and Terjanian[95] each proved the case n = 14, while Kapferer[91] and Breusch[93] 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.[96]

All proofs for specific exponents used Fermat's technique of infinite descent,[citation needed] either in its original form, or in the form of descent on elliptic curves or abelian varieties. The details and auxiliary arguments, however, were often ad hoc and tied to the individual exponent under consideration.[97] Since they became ever more complicated as p increased, it seemed unlikely that the general case of Fermat's Last Theorem could be proved by building upon the proofs for individual exponents.[97] Although some general results on Fermat's Last Theorem were published in the early 19th century by Niels Henrik Abel and Peter Barlow,[98][99] the first significant work on the general theorem was done by Sophie Germain.[100]
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.[101] 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.[101][102] 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.[103] 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.[104]
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) that 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.[105] This conjecture was proven in 1983 by Gerd Faltings,[106] 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.[107] By 1978, Samuel Wagstaff had extended this to all primes less than 125,000.[108] By 1993, Fermat's Last Theorem had been proven for all primes less than four million.[109]

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"[110]:211 Taniyama–Shimura-Weil conjecture, proposed around 1955—which many mathematicians believed would be near to impossible to prove,[110]:223 and 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 1994, 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.
Taniyama–Shimura–Weil conjecture
Main article: Modularity theorem

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 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.[110]:211–215

Even after gaining serious attention, the conjecture was seen by contemporary mathematicians as extraordinarily difficult or perhaps inaccessible to proof.[110]:203–205, 223, 226 For example, Wiles' ex-supervisor John Coates states that it seemed "impossible to actually prove",[110]: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]."[110]:223
Frey's equation / Ribet's theorem
Main articles: Frey curve and 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.[111] 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[112] and 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) that included Frey's equation – and this was widely believed inaccessible to proof by contemporary mathematicians.[110]: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
Main articles: Andrew Wiles and Wiles's proof of Fermat's Last Theorem

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.[113]

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.[110]:229–230 His initial study suggested proof by induction,[110]:230–232, 249–252 and he based his initial work and first significant breakthrough on Galois theory[110]: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.[110]:258–259 However, by the summer of 1991, Iwasawa theory also seemed to not be reaching the central issues in the problem.[110]:259–260[114] 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 that seemed "tailor made" for the inductive part of his proof.[110]: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 Wiles used seemed to work correctly.[110]:261–265[115]

By mid-May 1993 Wiles felt able to tell his wife he thought he had solved the proof of Fermat's Last Theorem,[110]: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.[116] 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),[117] who alerted Wiles on 23 August 1993.[118]

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.[110]: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.[119]

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.[120] On 24 October 1994, Wiles submitted two manuscripts, "Modular elliptic curves and Fermat's Last Theorem"[121] and "Ring theoretic properties of certain Hecke algebras",[122] the second of which was co-authored with Taylor and proved that certain conditions were met that 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 proven.[123][124][125] 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
Reciprocal Integers (Inverse Fermat Equation)

The equation a^{1/m} + b^{1/m} = c^{1/m} can be considered the "inverse" Fermat equation. All solutions of this equation were computed by Lenstra in 1992.[126] In the case in which the mth roots are required to be real and positive, all solutions are given by[127]

a=rs^m
b=rt^m
c=r(s+t)^m

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

For the Diophantine equation a^{n/m} + b^{n/m} = c^{n/m} with n not equal to 1, Bennett, Glass, and Székely proved in 2004 for n > 2 that if n and m are coprime, then there are integer solutions if and only if 6 divides m, and a^{1/m}, b^{1/m}, and c^{1/m} are different complex 6th roots of the same real number.[128]
Negative exponents
n = –1

All primitive (pairwise coprime) integer solutions to the optic equation a^{-1}+b^{-1}=c^{-1} can be written as[129]

a=mn+m^2,
b=mn+n^2,
c=mn

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.[130][131] All primitive solutions to a^{-2}+b^{-2}=d^{-2} are given by

a=(v^2-u^2)(v^2+u^2), \,
b=2uv(v^2+u^2), \,
d=2uv(v^2-u^2), \,

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

c=(v^2+u^2)^2, \,

so (a, b, c) is a Pythagorean triple.
Integer n < –2

There are no solutions in integers for a^n+b^n=c^n for integers n < –2. If there were, the equation could be multiplied through by a^{|n|}b^{|n|}c^{|n|} to obtain (bc)^{|n|}+(ac)^{|n|}=(ab)^{|n|}, which is impossible by Fermat's Last Theorem.
Base values other than positive integers

Fermat's last theorem can easily be extended to positive rationals:

\left(\frac{a}{x}\right)^n+\left(\frac{b}{y}\right)^n=\left(\frac{c}{z}\right)^n

can have no solutions with n > 2, because any solution could be rearranged as:

(ayz)^n+(bxz)^n=(cxy)^n,

to which Fermat's Last Theorem applies.
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.[132] 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.[133] Another prize was offered in 1883 by the Academy of Brussels.[134]

In 1908, the German industrialist and amateur mathematician Paul Wolfskehl bequeathed 100,000 gold marks—a large sum at the time—to the Göttingen Academy of Sciences to offer as a prize for a complete proof of Fermat's Last Theorem.[135] 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.[136] Wiles collected the Wolfskehl prize money, then worth $50,000, on 27 June 1997.[137]

Prior to Wiles's proof, thousands of incorrect proofs were submitted to the Wolfskehl committee, amounting to roughly 10 feet (3 meters) of correspondence.[138] 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".[139] 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."[134]
See also

Beal's conjecture
Diophantus II.VIII
Euler's sum of powers conjecture
Fermat's Last Theorem in fiction
Proof of impossibility
Sophie Germain prime
Sums of powers, a list of related conjectures and theorems
Wall–Sun–Sun prime

Footnotes

If the exponent "n" were not prime or 4, then it would be possible to write n either as a product of two smaller integers (n = P*Q) in which P is a prime number greater than 2, 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 that is smaller than N, as well; or else as n would be a power of 2 greater than four and writing n=4*Q, the same argument would hold.
For example, \left((j^r+1)^s\right)^r + \left(j(j^r+1)^s)\right)^r = (j^r+1)^{rs+1}.

This elliptic curve was first suggested in the 1960s by Yves Hellegouarch (de), 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

Stark, pp. 151–155.
Stillwell J (2003). Elements of Number Theory. New York: Springer-Verlag. pp. 110–112. ISBN 0-387-95587-9.
Aczel, pp. 13–15
Singh, pp. 18–20.
Singh, p. 6.
Stark, pp. 145–146.
Singh, pp. 50–51.
Stark, p. 145.
Aczel, pp. 44–45; Singh, pp. 56–58.
Aczel, pp. 14–15.
Stark, pp. 44–47.
Friberg, pp. 333– 334.
Dickson, p. 731; Singh, pp. 60–62; Aczel, p. 9.
T. Heath, Diophantus of Alexandria Second Edition, Cambridge University Press, 1910, reprinted by Dover, NY, 1964, pp. 144-145
Panchishkin, p. 341
Singh, pp. 62–66.
Dickson, p. 731.
Singh, p. 67; Aczel, p. 10.
Ribenboim, pp. 13, 24.
van der Poorten, Notes and Remarks 1.2, p. 5.
van der Poorten, loc. cit.
André Weil (1984). Number Theory: An approach through history. From Hammurapi to Legendre. Basel, Switzerland: Birkhäuser. p. 104.
BBC Documentary
Freeman L. "Fermat's One Proof". Retrieved 23 May 2009.
Dickson, pp. 615–616; Aczel, p. 44.
Ribenboim, pp. 15–24.
Frénicle de Bessy, Traité des Triangles Rectangles en Nombres, vol. I, 1676, Paris. Reprinted in Mém. Acad. Roy. Sci., 5, 1666–1699 (1729).
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).
Kausler CF (1802). "Nova demonstratio theorematis nec summam, nec differentiam duorum cuborum cubum esse posse". Novi Acta Acad. Petrop. 13: 245–253.
Barlow P (1811). An Elementary Investigation of Theory of Numbers. St. Paul's Church-Yard, London: J. Johnson. pp. 144–145.
Legendre AM (1830). Théorie des Nombres (Volume II) (3rd ed.). Paris: Firmin Didot Frères. Reprinted in 1955 by A. Blanchard (Paris).
Schopis (1825). Einige Sätze aus der unbestimmten Analytik. Gummbinnen: Programm.
Terquem O (1846). "Théorèmes sur les puissances des nombres". Nouv. Ann. Math. 5: 70–87.
Bertrand J (1851). Traité Élémentaire d'Algèbre. Paris: Hachette. pp. 217–230, 395.
Lebesgue VA (1853). "Résolution des équations biquadratiques z2 = x4 ± 2my4, z2 = 2mx4 − y4, 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.
Pepin T (1883). "Étude sur l'équation indéterminée ax4 + by4 = cz2". Atti Accad. Naz. Lincei 36: 34–70.
Tafelmacher WLA (1893). "Sobre la ecuación x4 + y4 = z4". Ann. Univ. Chile 84: 307–320.
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.
Bendz TR (1901). Öfver diophantiska ekvationen xn + yn = zn. Uppsala: Almqvist & Wiksells Boktrycken.
Gambioli D (1901). "Memoria bibliographica sull'ultimo teorema di Fermat". Period. Mat. 16: 145–192.
Kronecker L (1901). Vorlesungen über Zahlentheorie, vol. I. Leipzig: Teubner. pp. 35–38. Reprinted by New York:Springer-Verlag in 1978.
Bang A (1905). "Nyt Bevis for at Ligningen x4 − y4 = z4, ikke kan have rationale Løsinger". Nyt Tidsskrift Mat. 16B: 35–36.
Sommer J (1907). Vorlesungen über Zahlentheorie. Leipzig: Teubner.
Bottari A (1908). "Soluzione intere dell'equazione pitagorica e applicazione alla dimostrazione di alcune teoremi della teoria dei numeri". Period. Mat. 23: 104–110.
Rychlik K (1910). "On Fermat's last theorem for n = 4 and n = 3 (in Bohemian)". Časopis Pěst. Mat. 39: 65–86.
Nutzhorn F (1912). "Den ubestemte Ligning x4 + y4 = z4". Nyt Tidsskrift Mat. 23B: 33–38.
Carmichael RD (1913). "On the impossibility of certain Diophantine equations and systems of equations". Amer. Math. Monthly (Mathematical Association of America) 20 (7): 213–221. doi:10.2307/2974106. JSTOR 2974106.
Hancock H (1931). Foundations of the Theory of Algebraic Numbers, vol. I. New York: Macmillan.
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.
Grant, Mike, and Perella, Malcolm, "Descending to the irrational", Mathematical Gazette 83, July 1999, pp.263–267.
Barbara, Roy, "Fermat's last theorem in the case n=4", Mathematical Gazette 91, July 2007, 260–262.
Dolan, Stan, "Fermat's method of descente infinie", Mathematical Gazette 95, July 2011, 269–271.
Ribenboim, pp. 1–2.
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.
Euler L (1770) Vollständige Anleitung zur Algebra, Roy. Acad. Sci., St. Petersburg.
Freeman L. "Fermat's Last Theorem: Proof for n = 3". Retrieved 23 May 2009.
Ribenboim, pp. 24–25; Mordell, pp. 6–8; Edwards, pp. 39–40.
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.
Ribenboim, pp. 33, 37–41.
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.
Calzolari L (1855). Tentativo per dimostrare il teorema di Fermat sull'equazione indeterminata xn + yn = zn. Ferrara.
Lamé G (1865). "Étude des binômes cubiques x3 ± y3". C. R. Acad. Sci. Paris 61: 921–924, 961–965.
Tait PG (1872). "Mathematical Notes". Proc. Roy. Soc. Edinburgh 7: 144.
Günther S (1878). "Über die unbestimmte Gleichung x3 + y3 = z3". Sitzungsberichte Böhm. Ges. Wiss.: 112–120.
Krey H (1909). "Neuer Beweis eines arithmetischen Satzes". Math. Naturwiss. Blätter 6: 179–180.
Stockhaus H (1910). Beitrag zum Beweis des Fermatschen Satzes. Leipzig: Brandstetter.
Carmichael RD (1915). Diophantine Analysis. New York: Wiley.
van der Corput JG (1915). "Quelques formes quadratiques et quelques équations indéterminées". Nieuw Archief Wisk. 11: 45–75.
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.
Duarte FJ (1944). "Sobre la ecuación x3 + y3 + z3 = 0". Ciencias Fis. Mat. Naturales (Caracas) 8: 971–979.
Freeman L. "Fermat's Last Theorem: Proof for n = 5". Retrieved 23 May 2009.
Ribenboim, p. 49; Mordell, p. 8–9; Aczel, p. 44; Singh, p. 106.
Ribenboim, pp. 55–57.
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. Check date values in: |date= (help)
Lebesgue VA (1843). "Théorèmes nouveaux sur l'équation indéterminée x5 + y5 = az5". J. Math. Pures Appl. 8: 49–70.
Lamé G (1847). "Mémoire sur la résolution en nombres complexes de l'équation A5 + B5 + C5 = 0". J. Math. Pures Appl. 12: 137–171.
Gambioli D (1903/4). "Intorno all'ultimo teorema di Fermat". Il Pitagora 10: 11–13, 41–42. Check date values in: |date= (help)
Werebrusow AS (1905). "On the equation x5 + y5 = Az5 (in Russian)". Moskov. Math. Samml. 25: 466–473.
Rychlik K (1910). "On Fermat's last theorem for n = 5 (in Bohemian)". Časopis Pěst. Mat. 39: 185–195, 305–317.
Terjanian G (1987). "Sur une question de V. A. Lebesgue". Ann. Inst. Fourier 37: 19–37. doi:10.5802/aif.1096.
Ribenboim, pp. 57–63; Mordell, p. 8; Aczel, p. 44; Singh, p. 106.
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.
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.
Freeman L. "Fermat's Last Theorem: Proof for n = 7". Retrieved 23 May 2009.
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.
Pepin T (1876). "Impossibilité de l'équation x7 + y7 + z7 = 0". C. R. Acad. Sci. Paris 82: 676–679, 743–747.
Maillet E (1897). "Sur l'équation indéterminée axλt + byλt = czλt". Assoc. Française Avanc. Sci., St. Etienne (sér. II) 26: 156–168.
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).
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.
Lind B (1909). "Einige zahlentheoretische Sätze". Arch. Math. Phys. 15: 368–369.
Kapferer H (1913). "Beweis des Fermatschen Satzes für die Exponenten 6 und 10". Archiv Math. Phys. 21: 143–146.
Swift E (1914). "Solution to Problem 206". Amer. Math. Monthly 21: 238–239.
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.
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).
Terjanian G (1974). "L'équation x14 + y14 = z14 en nombres entiers". Bull. Sci. Math. (sér. 2) 98: 91–95.
Edwards, pp. 73–74.
Edwards, p. 74.
Dickson, p. 733.
Ribenboim P (1979). 13 Lectures on Fermat's Last Theorem. New York: Springer Verlag. pp. 51–54. ISBN 978-0-387-90432-0.
Singh, pp. 97–109.
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.
Aczel, p. 57.
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.
Adleman LM, Heath-Brown DR (June 1985). "The first case of Fermat's last theorem". Inventiones Mathematicae (Berlin: Springer) 79 (2): 409–416. doi:10.1007/BF01388981.
Aczel, pp. 84–88; Singh, pp. 232–234.
Faltings G (1983). "Endlichkeitssätze für abelsche Varietäten über Zahlkörpern". Inventiones Mathematicae 73 (3): 349–366. doi:10.1007/BF01388432.
Ribenboim P (1979). 13 Lectures on Fermat's Last Theorem. New York: Springer Verlag. p. 202. ISBN 978-0-387-90432-0.
Wagstaff SS, Jr. (1978). "The irregular primes to 125000". Math. Comp. (American Mathematical Society) 32 (142): 583–591. doi:10.2307/2006167. JSTOR 2006167. (PDF)
Buhler J, Crandell R, Ernvall R, Metsänkylä T (1993). "Irregular primes and cyclotomic invariants to four million". Math. Comp. (American Mathematical Society) 61 (203): 151–153. doi:10.2307/2152942. JSTOR 2152942.
[Fermat's Last Theorem, Simon Singh, 1997, ISBN 1-85702-521-0
Frey G (1986). "Links between stable elliptic curves and certain diophantine equations". Ann. Univ. Sarav. Ser. Math. 1: 1–40.
Singh, pp. 194–198; Aczel, pp. 109–114.
Singh, p. 205; Aczel, pp. 117–118.
Singh, pp. 237–238; Aczel, pp. 121–122.
Singh, pp. 239–243; Aczel, pp. 122–125.
Singh, pp. 244–253; Aczel, pp. 1–4, 126–128.
Aczel, pp. 128–130.
Singh, p. 257.
Singh, pp. 269–274.
Singh, pp. 275–277; Aczel, pp. 132–134.
Wiles, Andrew (1995). "Modular elliptic curves and Fermat's Last Theorem" (PDF). Annals of Mathematics 141 (3): 443–551. doi:10.2307/2118559. JSTOR 2118559. OCLC 37032255.
Taylor R, Wiles A (1995). "Ring theoretic properties of certain Hecke algebras". Annals of Mathematics 141 (3): 553–572. doi:10.2307/2118560. JSTOR 2118560. OCLC 37032255.
Diamond, Fred (1996). "On deformation rings and Hecke rings". Annals of Mathematics. Second Series 144 (1): 137–166. doi:10.2307/2118586. ISSN 0003-486X. MR 1405946.
Conrad, Brian; Diamond, Fred; Taylor, Richard (1999). "Modularity of certain potentially Barsotti-Tate Galois representations". Journal of the American Mathematical Society 12 (2): 521–567. doi:10.1090/S0894-0347-99-00287-8. ISSN 0894-0347. MR 1639612.
Breuil, Christophe; Conrad, Brian; Diamond, Fred; Taylor, Richard (2001). "On the modularity of elliptic curves over Q: wild 3-adic exercises". Journal of the American Mathematical Society 14 (4): 843–939. doi:10.1090/S0894-0347-01-00370-8. ISSN 0894-0347. MR 1839918.
Lenstra, Jr. H.W. (1992). On the inverse Fermat equation, Discrete Mathematics, 106–107, pp. 329–331.
Newton, M., "A radical diophantine equation", Journal of Number Theory 13 (1981), 495–498.
Bennett, Curtis D.; Glass, A. M. W.; Székely, Gábor J. (2004). "Fermat's last theorem for rational exponents". American Mathematical Monthly 111 (4): 322–329. doi:10.2307/4145241. MR 2057186.
Dickson, pp. 688–691
Voles, Roger (July 1999). "Integer solutions of a−2 + b−2 = d−2". Mathematical Gazette 83: 269–271.
Richinick, Jennifer (July 2008). "The upside-down Pythagorean Theorem". Mathematical Gazette 92: 313–317.
Aczel, p. 69; Singh, p. 105.
Aczel, p. 69.
Koshy T (2001). Elementary number theory with applications. New York: Academic Press. p. 544. ISBN 978-0-12-421171-1.
Singh, pp. 120–125, 131–133, 295–296; Aczel, p. 70.
Singh, pp. 120–125.
Singh, p. 284
Singh, p. 295.

Singh, pp. 295–296.

Bibliography

Aczel, Amir (30 September 1996). Fermat's Last Theorem: Unlocking the Secret of an Ancient Mathematical Problem. Four Walls Eight Windows. ISBN 978-1-56858-077-7.
Dickson LE (1919). History of the Theory of Numbers. Volume II. Diophantine Analysis. New York: Chelsea Publishing. pp. 545–550, 615–621, 688–691, 731–776.
Edwards, HM (1997). Fermat's Last Theorem. A Genetic Introduction to Algebraic Number Theory. Graduate Texts in Mathematics 50. New York: Springer-Verlag.
Friberg, Joran (2007). Amazing Traces of a Babylonian Origin in Greek Mathematics. World Scientific Publishing Company. ISBN 978-981-270-452-8.
Kleiner I (2000). "From Fermat to Wiles: Fermat's Last Theorem Becomes a Theorem" (PDF). Elem. Math. 55: 19–37. doi:10.1007/PL00000079.
Mordell LJ (1921). Three Lectures on Fermat's Last Theorem. Cambridge: Cambridge University Press.
Panchishkin, Alekseĭ Alekseevich (2007). Introduction to Modern Number Theory (Encyclopedia of Mathematical Sciences. Springer Berlin Heidelberg New York. ISBN 978-3-540-20364-3.
Ribenboim P (2000). Fermat's Last Theorem for Amateurs. New York: Springer-Verlag. ISBN 978-0-387-98508-4.
Singh S (October 1998). Fermat's Enigma. New York: Anchor Books. ISBN 978-0-385-49362-8.
Stark H (1978). An Introduction to Number Theory. MIT Press. ISBN 0-262-69060-8.

Further reading

Bell, Eric T. (6 August 1998) [1961]. The Last Problem. New York: The Mathematical Association of America. ISBN 978-0-88385-451-8.
Benson, Donald C. (5 April 2001). The Moment of Proof: Mathematical Epiphanies. Oxford University Press. ISBN 978-0-19-513919-8.
Brudner, Harvey J. (1994). Fermat and the Missing Numbers. WLC, Inc. ISBN 978-0-9644785-0-3.
Edwards, H. M. (March 1996) [1977]. Fermat's Last Theorem. New York: Springer-Verlag. ISBN 978-0-387-90230-2.
Faltings G (July 1995). "The Proof of Fermat's Last Theorem by R. Taylor and A. Wiles" (PDF). Notices of the AMS 42 (7): 743–746. ISSN 0002-9920.
Mozzochi, Charles (7 December 2000). The Fermat Diary. American Mathematical Society. ISBN 978-0-8218-2670-6.
Ribenboim P (1979). 13 Lectures on Fermat's Last Theorem. New York: Springer Verlag. ISBN 978-0-387-90432-0.
van der Poorten, Alf (6 March 1996). Notes on Fermat's Last Theorem. WileyBlackwell. ISBN 978-0-471-06261-5.
Saikia, Manjil P (July 2011). "A Study of Kummer's Proof of Fermat's Last Theorem for Regular Primes" (PDF). IISER Mohali (India) Summer Project Report.

External links
Wikibooks has more on the topic of: Fermat's Last Theorem

Wiles (1995). "Modular elliptic curves and Fermat's Last Theorem" (PDF). Retrieved 9 February 2013. Scientific article by Andrew Wiles
Daney, Charles (2003). "The Mathematics of Fermat's Last Theorem". Retrieved 5 August 2004.
The bluffer's guide to Fermat's Last Theorem
Elkies, Noam D. "Tables of Fermat "near-misses" — approximate solutions of xn + yn = zn".
Freeman, Larry (2005). "Fermat's Last Theorem Blog". Blog that covers the history of Fermat's Last Theorem from Fermat to Wiles.
Hazewinkel, Michiel, ed. (2001), "Fermat's last theorem", Encyclopedia of Mathematics, Springer, ISBN 978-1-55608-010-4
Ribet, Ken (1995). "Galois representations and modular forms" (PDF). Discusses various material that is related to the proof of Fermat's Last Theorem: elliptic curves, modular forms, Galois representations and their deformations, Frey's construction, and the conjectures of Serre and of Taniyama–Shimura.
Shay, David (2003). "Fermat's Last Theorem". Retrieved 5 August 2004. The story, the history and the mystery.
Weisstein, Eric W., "Fermat's Last Theorem", MathWorld.
O'Connor JJ, Robertson EF (1996). "Fermat's last theorem". Retrieved 5 August 2004.
"The Proof". The title of one edition of the PBS television series NOVA, discusses Andrew Wiles's effort to prove Fermat's Last Theorem.
"Documentary Movie on Fermat's Last Theorem (1996)". Simon Singh and John Lynch's film tells the story of Andrew Wiles.
Beal Fermat and Pythagora's Triplets (sic)

Mathematics Encyclopedia

Retrieved from "http://en.wikipedia.org/"
All text is available under the terms of the GNU Free Documentation License

Home - Hellenica World