Fermat's right triangle theorem

Last updated

Two right triangles with the two legs of the top one equal to the leg and hypotenuse of the bottom one. For these lengths,
a
2
{\displaystyle a^{2}}
,
b
2
{\displaystyle b^{2}}
, and
c
2
{\displaystyle c^{2}}
form an arithmetic progression separated by a gap of
d
2
{\displaystyle d^{2}}
. It is not possible for all four lengths
a
{\displaystyle a}
,
b
{\displaystyle b}
,
c
{\displaystyle c}
, and
d
{\displaystyle d}
to be integers. Fermat right triangles.svg
Two right triangles with the two legs of the top one equal to the leg and hypotenuse of the bottom one. For these lengths, , , and form an arithmetic progression separated by a gap of . It is not possible for all four lengths , , , and to be integers.

Fermat's right triangle theorem is a non-existence proof in number theory, published in 1670 among the works of Pierre de Fermat, soon after his death. It is the only complete proof given by Fermat. [1] It has many equivalent formulations, one of which was stated (but not proved) in 1225 by Fibonacci. In its geometric forms, it states:

Contents

More abstractly, as a result about Diophantine equations (integer or rational-number solutions to polynomial equations), it is equivalent to the statements that:

An immediate consequence of the last of these formulations is that Fermat's Last Theorem is true in the special case that its exponent is 4.

Formulation

Squares in arithmetic progression

In 1225, Emperor Frederick II challenged the mathematician Fibonacci to take part in a mathematical contest against several other mathematicians, with three problems set by his court philosopher John of Palermo. The first of these problems asked for three rational numbers whose squares were equally spaced five units apart, solved by Fibonacci with the three numbers , , and . In The Book of Squares , published later the same year by Fibonacci, he solved the more general problem of finding triples of square numbers that are equally spaced from each other, forming an arithmetic progression. Fibonacci called the gap between these numbers a congruum. [2] One way of describing Fibonacci's solution is that the numbers to be squared are the difference of legs, hypotenuse, and sum of legs of a Pythagorean triangle, and that the congruum is four times the area of the same triangle. [3] Fibonacci observed that it is impossible for a congruum to be a square number itself, but did not present a satisfactory proof of this fact. [4]

If three squares , , and could form an arithmetic progression whose congruum was also a square , then these numbers would satisfy the Diophantine equations

That is, by the Pythagorean theorem, they would form two integer-sided right triangles in which the pair gives one leg and the hypotenuse of the smaller triangle and the same pair also forms the two legs of the larger triangle. But if (as Fibonacci asserted) no square congruum can exist, then there can be no two integer right triangles that share two sides in this way. [5]

Areas of right triangles

Because the congrua are exactly the numbers that are four times the area of a Pythagorean triangle, and multiplication by four does not change whether a number is square, the existence of a square congruum is equivalent to the existence of a Pythagorean triangle with a square area. It is this variant of the problem that Fermat's proof concerns: he shows that there is no such triangle. In considering this problem, Fermat was inspired not by Fibonacci but by an edition of Arithmetica by Diophantus, published in a translation into French in 1621 by Claude Gaspar Bachet de Méziriac. [6] This book described various special right triangles whose areas had forms related to squares, but did not consider the case of areas that were themselves square. [7]

By rearranging the equations for the two Pythagorean triangles above, and then multiplying them together, one obtains the single Diophantine equation

which can be simplified by introducing a new variable to

Conversely, any three positive integers obeying the equation lead to a square congruum: for these numbers, the squares , , and form an arithmetic progression with congruum , which is a square itself. Thus, the solvability of is equivalent to the existence of a square congruum. But, if Fermat's Last Theorem had a counterexample for the exponent , an integer solution to the equation , then squaring one of the three numbers in the counterexample would give three numbers that solve the equation . Therefore, Fermat's proof that no Pythagorean triangle has a square area implies the truth of the exponent- case of Fermat's Last Theorem. [7]

Another equivalent formulation of the same problem involves congruent numbers, the numbers that are areas of right triangles whose three sides are all rational numbers. By multiplying the sides by a common denominator, any congruent number may be transformed into the area of a Pythagorean triangle, from which it follows that the congruent numbers are exactly the numbers formed by multiplying a congruum by the square of a rational number. [8] Therefore, the existence of a square congruum is equivalent to the statement that the number 1 is not a congruent number. [9] Another more geometric way of stating this formulation is that it is impossible for a square (the geometric shape) and a right triangle to have both equal areas and all sides commensurate with each other. [10]

Elliptic curve

The elliptic curve y = x(x + 1)(x - 1). The three rational points (-1,0), (0,0), and (1,0) are the crossings of this curve with the x-axis. Elliptic curve y^2 = x^3 - x.svg
The elliptic curve y = x(x + 1)(x 1). The three rational points (1,0), (0,0), and (1,0) are the crossings of this curve with the x-axis.

Yet another equivalent form of Fermat's theorem involves the elliptic curve consisting of the points whose Cartesian coordinates satisfy the equation

The points (1,0), (0,0), and (1,0), provide obvious solutions to this equation. Fermat's theorem is equivalent to the statement that these are the only points on the curve for which both and are rational. More generally, the right triangles with rational sides and area correspond one-for-one with the rational points with positive -coordinate on the elliptic curve . [11]

Fermat's proof

During his lifetime, Fermat challenged several other mathematicians to prove the non-existence of a Pythagorean triangle with square area, but did not publish the proof himself. However, he wrote a proof in his copy of Diophantus's Arithmetica, the same copy in which he wrote that he could prove Fermat's Last Theorem. Fermat's son Clement-Samuel published an edition of this book, including Fermat's marginal notes with the proof of the right triangle theorem, in 1670. [12]

Fermat's proof is a proof by infinite descent. It shows that, from any example of a Pythagorean triangle with square area, one can derive a smaller example. Since Pythagorean triangles have positive integer areas, and there does not exist an infinite descending sequence of positive integers, there also cannot exist a Pythagorean triangle with square area. [13]

In more detail, suppose that , , and are the integer sides of a right triangle with square area. By dividing by any common factors, one can assume that this triangle is primitive [10] and from the known form of all primitive Pythagorean triples, one can set , , and , by which the problem is transformed into finding relatively prime integers and (one of which is even) such that the area is square. For this number to be a square, its four linear factors , , , and (which are relatively prime) must themselves be squares; let and . Both and must be odd since exactly one of or is even and the other is odd. Therefore, both and are even, and one of them is divisible by 4. Dividing them by two produces two more integers and , one of which is even by the previous sentence. Because is a square, and are the legs of another primitive Pythagorean triangle whose area is . Since is itself a square and since is even, is a square. Thus, any Pythagorean triangle with square area leads to a smaller Pythagorean triangle with square area, completing the proof. [14]

Notes

  1. Edwards (2000). Many subsequent mathematicians published proofs, including Gottfried Wilhelm Leibniz (1678), Leonhard Euler (1747), and Bernard Frenicle de Bessy (before 1765); see Dickson (1920) and Goldstein (1995).
  2. Bradley (2006).
  3. Beiler (1964).
  4. Ore (2012); Dickson (1920).
  5. The fact that there can be no two right triangles that share two of their sides, and the connection between this problem and the problem of squares in arithmetic progression, is described as "well known" by Cooper & Poirel (2008)
  6. Edwards (2000).
  7. 1 2 Stillwell (1998).
  8. Conrad (2008); Koblitz (1993 , p. 3).
  9. Conrad (2008), Theorem 2; Koblitz (1993), Exercise 3, p. 5.
  10. 1 2 Dickson (1920).
  11. Koblitz (1993), Proposition 19, pp. 46–47; Kato & Saitō (2000).
  12. Edwards (2000); Dickson (1920). For other proofs, see Grant & Perella (1999) and Barbara (2007).
  13. Edwards (2000); Dickson (1920).
  14. Edwards (2000); Dickson (1920); Stillwell (1998).

Related Research Articles

<span class="mw-page-title-main">Diophantine equation</span> Polynomial equation whose integer solutions are sought

In mathematics, a Diophantine equation is an equation, typically a polynomial equation in two or more unknowns with integer coefficients, for which only integer solutions are of interest. A linear Diophantine equation equates to a constant the sum of two or more monomials, each of degree one. An exponential Diophantine equation is one in which unknowns can appear in exponents.

<span class="mw-page-title-main">Number theory</span> Mathematics of integer properties

Number theory is a branch of pure mathematics devoted primarily to the study of the integers and arithmetic functions. German mathematician Carl Friedrich Gauss (1777–1855) said, "Mathematics is the queen of the sciences—and number theory is the queen of mathematics." Number theorists study prime numbers as well as the properties of mathematical objects constructed from integers, or defined as generalizations of the integers.

<span class="mw-page-title-main">Pythagorean triple</span> Integer side lengths of a right triangle

A Pythagorean triple consists of three positive integers a, b, and c, such that a2 + b2 = c2. Such a triple is commonly written (a, b, c), a well-known example is (3, 4, 5). If (a, b, c) is a Pythagorean triple, then so is (ka, kb, kc) for any positive integer k. A triangle whose side lengths are a Pythagorean triple is a right triangle and called a Pythagorean triangle.

In mathematics, a quadratic irrational number is an irrational number that is the solution to some quadratic equation with rational coefficients which is irreducible over the rational numbers. Since fractions in the coefficients of a quadratic equation can be cleared by multiplying both sides by their least common denominator, a quadratic irrational is an irrational root of some quadratic equation with integer coefficients. The quadratic irrational numbers, a subset of the complex numbers, are algebraic numbers of degree 2, and can therefore be expressed as

In mathematics, a proof by infinite descent, also known as Fermat's method of descent, is a particular kind of proof by contradiction used to show that a statement cannot possibly hold for any number, by showing that if the statement were to hold for a number, then the same would be true for a smaller number, leading to an infinite descent and ultimately a contradiction. It is a method which relies on the well-ordering principle, and is often used to show that a given equation, such as a Diophantine equation, has no solutions.

In geometry, a Heronian triangle is a triangle whose side lengths a, b, and c and area A are all positive integers. Heronian triangles are named after Heron of Alexandria, based on their relation to Heron's formula which Heron demonstrated with the example triangle of sides 13, 14, 15 and area 84.

In additive number theory, Fermat's theorem on sums of two squares states that an odd prime p can be expressed as:

<span class="mw-page-title-main">Special right triangle</span> Right triangle with a feature making calculations on the triangle easier

A special right triangle is a right triangle with some regular feature that makes calculations on the triangle easier, or for which simple formulas exist. For example, a right triangle may have angles that form simple relationships, such as 45°–45°–90°. This is called an "angle-based" right triangle. A "side-based" right triangle is one in which the lengths of the sides form ratios of whole numbers, such as 3 : 4 : 5, or of other special numbers such as the golden ratio. Knowing the relationships of the angles or ratios of sides of these special right triangles allows one to quickly calculate various lengths in geometric problems without resorting to more advanced methods.

<span class="mw-page-title-main">Congruent number</span>

In number theory, a congruent number is a positive integer that is the area of a right triangle with three rational number sides. A more general definition includes all positive rational numbers with this property.

<span class="mw-page-title-main">Pythagorean prime</span>

A Pythagorean prime is a prime number of the form . Pythagorean primes are exactly the odd prime numbers that are the sum of two squares; this characterization is Fermat's theorem on sums of two squares.

<span class="mw-page-title-main">Congruum</span> Spacing between equally-spaced square numbers

In number theory, a congruum is the difference between successive square numbers in an arithmetic progression of three squares. That is, if , , and are three square numbers that are equally spaced apart from each other, then the spacing between them, , is called a congruum.

<span class="mw-page-title-main">Pythagorean quadruple</span> Four integers where the sum of the squares of three equals the square of the fourth

A Pythagorean quadruple is a tuple of integers a, b, c, and d, such that a2 + b2 + c2 = d2. They are solutions of a Diophantine equation and often only positive integer values are considered. However, to provide a more complete geometric interpretation, the integer values can be allowed to be negative and zero (thus allowing Pythagorean triples to be included) with the only condition being that d > 0. In this setting, a Pythagorean quadruple (a, b, c, d) defines a cuboid with integer side lengths |a|, |b|, and |c|, whose space diagonal has integer length d; with this interpretation, Pythagorean quadruples are thus also called Pythagorean boxes. In this article we will assume, unless otherwise stated, that the values of a Pythagorean quadruple are all positive integers.

In mathematics and statistics, sums of powers occur in a number of contexts:

<span class="mw-page-title-main">Fermat's Last Theorem</span> 17th-century conjecture proved by Andrew Wiles in 1994

In number theory, Fermat's Last Theorem states that no three positive integers a, b, and c satisfy the equation an + bn = cn for any integer value of n greater than 2. The cases n = 1 and n = 2 have been known since antiquity to have infinitely many solutions.

<span class="mw-page-title-main">Irrational number</span> Number that is not a ratio of integers

In mathematics, the irrational numbers are all the real numbers that are not rational numbers. That is, irrational numbers cannot be expressed as the ratio of two integers. When the ratio of lengths of two line segments is an irrational number, the line segments are also described as being incommensurable, meaning that they share no "measure" in common, that is, there is no length, no matter how short, that could be used to express the lengths of both of the two given segments as integer multiples of itself.

Besides Euclid's formula, many other formulas for generating Pythagorean triples have been developed.

Fermat's Last Theorem is a theorem in number theory, originally stated by Pierre de Fermat in 1637 and proven by Andrew Wiles in 1995. The statement of the theorem involves an integer exponent n larger than 2. In the centuries following the initial statement of the result and before its general proof, various proofs were devised for particular values of the exponent n. Several of these proofs are described below, including Fermat's proof in the case n = 4, which is an early example of the method of infinite descent.

<span class="mw-page-title-main">Pythagorean theorem</span> Relation between sides of a right triangle

In mathematics, the Pythagorean theorem or Pythagoras' theorem is a fundamental relation in Euclidean geometry between the three sides of a right triangle. It states that the area of the square whose side is the hypotenuse is equal to the sum of the areas of the squares on the other two sides.

<span class="mw-page-title-main">Integer triangle</span> Triangle with integer side lengths

An integer triangle or integral triangle is a triangle all of whose side lengths are integers. A rational triangle is one whose side lengths are rational numbers; any rational triangle can be rescaled by the lowest common denominator of the sides to obtain a similar integer triangle, so there is a close relationship between integer triangles and rational triangles.

References