Proofs of Fermat's theorem on sums of two squares

Proofs of Fermat's theorem on sums of two squares

Fermat's theorem on sums of two squares asserts that an odd prime number "p" can be expressed as

: p = x^2 + y^2

with integer "x" and "y" if and only if "p" is congruent to 1 (mod 4). The statement was announced by Fermat in 1640, but he supplied no proof.

The "only if" clause is easy: a perfect square is congruent to 0 or 1 modulo 4, hence a sum of two squares is congruent to 0, 1, or 2. An odd prime number is congruent to either 1 or 3 modulo 4, and the second possibility has just been ruled out. The first proof that such a representation exists was given by Leonhard Euler in 1747 and was quite complicated. Since then, many different proofs have been found. Among them, the proof using Minkowski's theorem about convex sets and Don Zagier's stunningly short proof based on involutions especially stand out.

Euler's proof by infinite descent

Euler succeeded in proving Fermat's theorem on sums of two squares in 1747, when he was forty years old. He communicated this in a letter to Goldbach dated 6 May 1747. The proof relies on infinite descent, and proceeds in five steps; the fifth step below is from another letter to Goldbach written in 1749, as the first letter was vague on that final step:

1. "The product of two numbers, each of which is a sum of two squares, is itself a sum of two squares."

::This is simply a restatement of the Brahmagupta-Fibonacci identity.

2. "If a number which is a sum of two squares is divisible by a prime which is a sum of two squares, then the quotient is a sum of two squares."

::Indeed, suppose for example that a^2 + b^2 is divisible by p^2+q^2 and that this latter is a prime. Then p^2 + q^2 divides

:::(pb-aq)(pb+aq) = p^2b^2 - a^2q^2 = p^2(a^2+b^2) - a^2(p^2+q^2).

::Since p^2+q^2 is a prime, it divides one of the two factors. Suppose that it divides pb-aq. Since

:::(a^2+b^2)(p^2+q^2) = (ap+bq)^2 + (aq-bp)^2,

::(Brahmagupta-Fibonacci identity) it follows that p^2+q^2 must divide (ap+bq)^2. So the equation can be divided by the square of p^2+q^2. Dividing the expression by (p^2+q^2)^2 yields:

:::frac{a^2+b^2}{p^2+q^2} = left(frac{ap+bq}{p^2+q^2} ight)^2 + left(frac{aq-bp}{p^2+q^2} ight)^2

::and thus expresses the quotient as a sum of two squares, as claimed.

::If p^2+q^2 divides pb+aq, a similar argument holds by using

:::(a^2+b^2)(q^2+p^2) = (aq+bp)^2 + (ap-bq)^2,

::(Brahmagupta-Fibonacci identity).

3. "If a number which can be written as a sum of two squares is divisible by a number which is not a sum of two squares, then the quotient has a factor which is not a sum of two squares."

::Suppose x divides a^2+b^2 and that the quotient, factored into its prime factors is p_1p_2cdots p_n. Then a^2+b^2 = x p_1p_2cdots p_n. If all factors p_i can be written as sums of two squares, then we can divide a^2+b^2 successively by p_1, p_2, etc., and applying the previous step we deduce that each quotient is a sum of two squares. This until we get to x, concluding that x would have to be the sum of two squares. So, by contraposition, if x is not the sum of two squares, then at least one of the primes p_i is not the sum of two squares.

4. "If a and b are relatively prime then every factor of a^2 + b^2 is a sum of two squares."

::This is the step that uses infinite descent. Let x be a factor of a^2+b^2. We can write :::a = mx pm c,qquad b = nx pm d::where c and d are at most half of x in absolute value. This gives::::a^2 + b^2 = m^2x^2pm 2mxc + c^2 + n^2x^2 pm 2nxd + d^2 = Ax + (c^2+d^2).::Therefore, c^2+d^2 must be divisible by x, say c^2+d^2 = yx. If c and d are not relatively prime, then their gcd cannot divide x (if it did, then it would divide a and b which we assume are relatively prime). Thus the square of the gcd divides y (as it divides c^2+d^2), giving us an expression of the form e^2+f^2 = zx for relatively prime e and f, and with z no more than half of x, since

:::zx = e^2 + f^2 leq c^2+d^2 leq left(frac{x}{2} ight)^2 + left(frac{x}{2} ight)^2 = frac{1}{2}x^2.

::If c and d are relatively prime, then we can use them directly instead of switching to e and f.

::If x is not the sum of two squares, then by the third step there must be a factor of z which is not the sum of two squares; call it w. This gives an infinite descent, going from x to a smaller number w, both not the sums of two squares but dividing a sum of two squares. Since an infinite descent is impossible, we conclude that x must be expressible as a sum of two squares, as claimed.

5. "Every prime of the form 4n+1 is a sum of two squares."

::If p=4n+1, then by Fermat's Little Theorem each of the numbers 1, 2^{4n}, 3^{4n},dots, (4n)^{4n} is congruent to one modulo p. The differences 2^{4n}-1, 3^{4n}-2^{4n},dots,(4n)^{4n}-(4n-1)^{4n} are therefore all divisible by p. Each of these differences can be factored as:::a^{4n}-b^{4n} = left(a^{2n}+b^{2n} ight)left(a^{2n}-b^{2n} ight).::Since p is prime, it must divide one of the two factors. If in any of the 4n-1 cases it divides the first factor, then by the previous step we conclude that p is itself a sum of two squares (since a and b differ by 1, they are relatively prime). So it is enough to show that p cannot always divide the second factor. If it divides all 4n-1 differences 2^{2n}-1, 3^{2n}-2^{2n},dots,(4n)^{2n}-(4n-1)^{2n}, then it would divide all 4n-2 differences of successive terms, all 4n-3 differences of the differences, and so forth. Since the kth differences of the sequence 1^k, 2^k, 3^k,dots are all equal to k!, the 2nth differences would all be constant and equal to (2n)!, which is certainly not divisible by p. Therefore, p cannot divide all the second factors which proves that p is indeed the sum of two squares.

Lagrange's proof through quadratic forms

Lagrange gave a proof in 1770 based on his general theory of integral quadratic forms. The following is a slight simplification of his argument, due to Gauss, which appears in article 182 of the Disquisitiones Arithmeticae.

A (binary) quadratic form will be taken to be an expression of the form ax^2 + 2bxy + cy^2 with a,b,c integers. A number n is said to be "represented by the form" if there exist integers x,y such that n = ax^2 + 2bxy + cy^2. Fermat's theorem on sums of two squares is then equivalent to the statement that a prime p is represented by the form x^2 + y^2 (i.e., a=c=1, b=0) exactly when p is congruent to 1 modulo 4.

The discriminant of the quadratic form is defined to be b^2 - ac (this is the definition due to Gauss; Lagrange did not require the xy term to have even coefficient, and defined the discriminant as b^2-4ac). The discriminant of x^2 + y^2 is then equal to -1.

Two forms ax^2 + 2bxy + cy^2 and rx'^2 + 2sx'y' + ty'^2 are "equivalent" if and only if there exist substitutions with integer coefficients: x = alpha x' + eta y': y = gamma x' + delta y'with alphadelta - etagamma = pm 1 such that, when substituted into the first form, yield the second. Equivalent forms are readily seen to have the same discriminant. Moreover, it is clear that equivalent forms will represent exactly the same integers.

Lagrange proved that all forms of discriminant -1 are equivalent. Thus, to prove Fermat's theorem it is enough to find "any" form of discriminant -1 that represents p. To do this, it suffices to find an integer m such that p divides m^2+1. For, finding such an integer, we can consider the form: px^2 + 2mxy + left(frac{m^2+1}{p} ight)y^2,which has discriminant -1 and represents p by setting x=1 and y=0.

Suppose then that p=4n+1. Again we invoke Fermat's Little Theorem: for any z relatively prime to p, we know that p divides z^{p-1} -1 = z^{4n}-1 = (z^{2n}-1)(z^{2n}+1). Moreover, by a theorem of Lagrange, the number of solutions modulo p to a congruence of degree q modulo p is at most q (this follows since the integers modulo p form a field, and a polynomial of degree q has at most q roots). So the congruence z^{2n} - 1 equiv 0 pmod{p} has at most 2n solutions among the numbers 1, 2, dots, p-1=4n. Therefore, there exists some positive integer z strictly smaller than p (and so relatively prime to p) such that p does not divide z^{2n}-1. Since p divides z^{4n}-1 = (z^{2n}-1)(z^{2n}+1), p must divide z^{2n}+1. Setting m=z^n completes the proof.

Dedekind's two proofs using Gaussian integers

Dedekind gave at least two proofs of Fermat's theorem on sums of two squares, both using the arithmetical properties of the Gaussian integers. One appears in section 27 of his exposition of ideals published in 1877; the second appeared in Supplement XI to Dirichlet's "Lectures on Number Theory", and was published in 1894.

1. First proof. If p is a positive odd rational prime, then we have i^p = (-1)^{frac{p-1}{2i, where i is the imaginary square root of -1. Consequently, if omega = x + iy is any Gaussian integer, then:omega^p = (x+yi)^p equiv x + (-1)^{frac{p-1}{2yi pmod{p}.If p is congruent to 1 modulo 4, then each Gaussian integer omega satisfies the congruence :omega^p equiv omega pmod{p},and therefore the ideal (p) is the product of two different prime ideals of degree 1 in Z [i] . Since Z [i] is a principal ideal domain, (in fact, a Euclidean domain), every ideal is principal, generated by an element of minimal norm. So if alpha is a generator of one of the ideal factors of (p), then we must have p = N(alpha) = N(a+bi) = a^2 + b^2 which gives Fermat's theorem.

2. Second proof. This proof builds on Lagrange's result that if p=4n+1, then there must be a number of the form m^2 + 1 which is a multiple of p. Since p does not divide the Gaussian factors m + i nor m-i (the quotients are clearly equal to egin{matrix}frac{m}{p}pm frac{i}{p}end{matrix}, which are not Gaussian integers), despite dividing the product, it follows that p cannot be a prime in Z [i] . Since it is not a prime, we must have a factorization in the Gaussian integers of the form p = (x+yi)(x-yi) for some integers x and y, which immediately yields that p = x^2 + y^2.

Zagier's "one-sentence proof"

Let "p" = 4"k" + 1 be prime. The set "S" = {("x", "y", "z") ∈ N3: "x"2 + 4"yz" = "p"} is finite and has two involutions, an obvious one ("x", "y", "z") → ("x", "z", "y"), whose fixed points correpond to representations of "p" as a sum of two squares, and a more complicated one,

: (x,y,z)mapstoegin{cases}(x+2z, z, y-x-z),quad extrm{if},,, x < y-z \ (2y-x, y, x-y+z),quad extrm{if},,, y-z < x < 2y\ (x-2y, x-y+z, y),quad extrm{if},,, x > 2yend{cases}

which has exactly one fixed point, (1, 1, "k"). However, the number of fixed points of an involution of a finite set "S" has the same parity as the cardinality of "S", therefore, this number is odd for the first involution as well, proving that "p" is a sum of two squares.

This proof, due to Zagier, is a simplification of an earlier proof by Heath-Brown, which in turn was inspired by a proof of Liouville. The technique of the proof is a combinatorial analogue of the topological principle that the Euler characteristics of a topological space with an involution and of its fixed point set have the same parity and is reminiscent of the use of "sign-reversing involutions" in the proofs of combinatorial bijections.

References

*Richard Dedekind, "The theory of algebraic integers".
*Harold M. Edwards, "Fermat's Last Theorem. A genetic introduction to algebraic number theory". Graduate Texts in Mathematics no. 50, Springer-Verlag, NY, 1977.
*C. F. Gauss, "Disquisitiones Arithmeticae" (English Edition). Transl. by Arthur A. Clarke. Springer-Verlag, 1986.
*D. R. Heath-Brown, "Fermat's two squares theorem". Invariant, 11 (1984) pp. 3-5.
*John Stillwell, Introduction to "Theory of Algebraic Integers" by Richard Dedekind. Cambridge Mathematical Library, Cambridge University Press, 1996.
* Don Zagier, "A one-sentence proof that every prime p &equiv; 1 mod 4 is a sum of two squares". Amer. Math. Monthly 97 (1990), no. 2, 144, doi|id=10.2307/2323918

External links

* [http://planetmath.org/encyclopedia/ProofOfThuesLemma.html Two more proofs at PlanetMath.org]
* [http://www.math.unh.edu/~dvf/532/Zagier A one-sentence proof of the theorem]
* [http://eprints.maths.ox.ac.uk/677/ reprint of Heath-Brown's proof, with commentary]


Wikimedia Foundation. 2010.

Игры ⚽ Поможем написать курсовую

Look at other dictionaries:

  • Fermat's theorem on sums of two squares — In number theory, Pierre de Fermat s theorem on sums of two squares states that an odd prime p is expressible as:p = x^2 + y^2,,with x and y integers, if and only if:p equiv 1 pmod{4}.The theorem is also known as Thue s Lemma, after Axel Thue.For …   Wikipedia

  • Sum of two squares — In mathematics, sums of two squares occur in a number of contexts:* The Pythagorean theorem says that the square on the hypotenuse of a right triangle is equal in area to the sum of the squares on the legs * Brahmagupta–Fibonacci identity says… …   Wikipedia

  • List of number theory topics — This is a list of number theory topics, by Wikipedia page. See also List of recreational number theory topics Topics in cryptography Contents 1 Factors 2 Fractions 3 Modular arithmetic …   Wikipedia

  • List of mathematics articles (P) — NOTOC P P = NP problem P adic analysis P adic number P adic order P compact group P group P² irreducible P Laplacian P matrix P rep P value P vector P y method Pacific Journal of Mathematics Package merge algorithm Packed storage matrix Packing… …   Wikipedia

  • Gaussian integer — In number theory, a Gaussian integer is a complex number whose real and imaginary part are both integers. The Gaussian integers, with ordinary addition and multiplication of complex numbers, form an integral domain, usually written as Z[i]. The… …   Wikipedia

  • Prime number — Prime redirects here. For other uses, see Prime (disambiguation). A prime number (or a prime) is a natural number greater than 1 that has no positive divisors other than 1 and itself. A natural number greater than 1 that is not a prime number is… …   Wikipedia

  • Pythagorean theorem — See also: Pythagorean trigonometric identity The Pythagorean theorem: The sum of the areas of the two squares on the legs (a and b) equals the area of the square on the hypotenuse (c) …   Wikipedia

  • Leonhard Euler — Infobox Scientist name = Leonhard Euler|box width = 300px |200px image width = 200px caption = Portrait by Johann Georg Brucker birth date = birth date|df=yes|1707|4|15 birth place = Basel, Switzerland death date = 18 September (O.S 7 September)… …   Wikipedia

  • Contributions of Leonhard Euler to mathematics — The 18th century Swiss mathematician Leonhard Euler (1707–1783) is among the most prolific and successful mathematicians in the history of the field. His seminal work had a profound impact in numerous areas of mathematics and he is widely… …   Wikipedia

  • mathematics — /math euh mat iks/, n. 1. (used with a sing. v.) the systematic treatment of magnitude, relationships between figures and forms, and relations between quantities expressed symbolically. 2. (used with a sing. or pl. v.) mathematical procedures,… …   Universalium

Share the article and excerpts

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