Trisha Shetty (Editor)

Quadratic reciprocity

Updated on
Edit
Like
Comment
Share on FacebookTweet on TwitterShare on LinkedInShare on Reddit
Quadratic reciprocity

In number theory, the law of quadratic reciprocity is a theorem about modular arithmetic that gives conditions for the solvability of quadratic equations modulo prime numbers. There are a number of equivalent statements of the theorem. One version of the law states that for p and q odd prime numbers,

Contents

( p q ) ( q p ) = ( 1 ) p 1 2 q 1 2

where

( p q )

denotes the Legendre symbol.

This law, combined with the properties of the Legendre symbol, means that any Legendre symbol can be calculated. This makes it possible to determine, for any quadratic equation, x 2 a ( mod p ) , where p is an odd prime, whether it has a solution. However, it does not provide any help at all for actually finding the solution. The solution can be found using quadratic residues.

The theorem was conjectured by Euler and Legendre and first proved by Gauss. He refers to it as the "fundamental theorem" in the Disquisitiones Arithmeticae and his papers, writing

The fundamental theorem must certainly be regarded as one of the most elegant of its type. (Art. 151)

Privately he referred to it as the "golden theorem." He published six proofs, and two more were found in his posthumous papers. There are now over 200 published proofs.

The first section of this article gives a special case of quadratic reciprocity that is representative of the general case. The second section gives the formulations of quadratic reciprocity found by Legendre and Gauss.

Count the number of proofs to the law of quadratic reciprocity given thus far in this book and devise another one.

Motivating example

Consider the polynomial f ( n ) = n 2 5 and its values for n N . The prime factorizations of these values are given as follows:

The prime numbers that appear as factors are 2,5, and all the prime numbers whose final digit is 1 or 9. No primes ending in 3 or 7 ever appear. Another way of phrasing this is that the primes p for which there exists an n such that n2 ≡ 5 (mod p) are precisely 2, 5, and those primes p that are ≡ 1 or 4 (mod 5). Or in other words, when p is a prime that is neither 2 nor 5, 5 is a quadratic residue modulo p iff p is 1 or 4 modulo 5. In other words, 5 is a quadratic residue modulo p iff p is a quadratic residue modulo 5.

The law of quadratic reciprocity gives a similar characterization of prime divisors of f(n) = n2c for any integer c.

Terminology, data, and two statements of the theorem

A quadratic residue (mod n) is any number congruent to a square (mod n). A quadratic nonresidue (mod n) is any number that is not congruent to a square (mod n). The adjective "quadratic" can be dropped if the context makes it clear that it is implied. When working modulo primes (as in this article), it is usual to treat zero as a special case. By doing so, the following statements become true:

  • Modulo a prime, there are an equal number of quadratic residues and nonresidues.
  • Modulo a prime, the product of two quadratic residues is a residue, the product of a residue and a nonresidue is a nonresidue, and the product of two nonresidues is a residue.
  • Table of quadratic residues

    This table is complete for odd primes less than 50. To check whether a number m is a quadratic residue mod one of these primes p, find am (mod p) and 0 ≤ a < p. If a is in row p, then m is a residue (mod p); if a is not in row p of the table, then m is a nonresidue (mod p).

    The quadratic reciprocity law is the statement that certain patterns found in the table are true in general.

    In this article, p and q always refer to distinct positive odd prime numbers.

    ±1 and the first supplement

    Trivially 1 is a quadratic residue for all primes. The question becomes more interesting for −1. Examining the table, we find −1 in rows 5, 13, 17, 29, 37, and 41 but not in rows 3, 7, 11, 19, 23, 31, 43 or 47. The former set of primes are all congruent to 1 modulo 4, and the latter are congruent to 3 modulo 4.

    First Supplement to Quadratic Reciprocity. The congruence x 2 1 ( mod p ) is solvable if and only if p is congruent to 1 modulo 4.

    ±2 and the second supplement

    Examining the table, we find 2 in rows 7, 17, 23, 31, 41, and 47, but not in rows 3, 5, 11, 13, 19, 29, 37, or 43. The former primes are all ≡ ±1 (mod 8), and the latter are all ≡ ±3 (mod 8). This leads to

    Second Supplement to Quadratic Reciprocity. The congruence x 2 2 ( mod p ) is solvable if and only if p is congruent to ±1 modulo 8.

    −2 is in rows 3, 11, 17, 19, 41, 43, but not in rows 5, 7, 13, 23, 29, 31, 37, or 47. The former are ≡ 1 or ≡ 3 (mod 8), and the latter are ≡ 5, 7 (mod 8).

    ±3

    3 is in rows 11, 13, 23, 37, and 47, but not in rows 5, 7, 17, 19, 29, 31, 41, or 43. The former are ≡ ±1 (mod 12) and the latter are all ≡ ±5 (mod 12).

    −3 is in rows 7, 13, 19, 31, 37, and 43 but not in rows 5, 11, 17, 23, 29, 41, or 47. The former are ≡ 1 (mod 3) and the latter ≡ 2 (mod 3).

    Since the only residue (mod 3) is 1, we see that −3 is a quadratic residue modulo every prime which is a residue modulo 3.

    ±5

    5 is in rows 11, 19, 29, 31, and 41 but not in rows 3, 7, 13, 17, 23, 37, 43, or 47. The former are ≡ ±1 (mod 5) and the latter are ≡ ±2 (mod 5).

    Since the only residues (mod 5) are ±1, we see that 5 is a quadratic residue modulo every prime which is a residue modulo 5.

    −5 is in rows 3, 7, 23, 29, 41, 43, and 47 but not in rows 11, 13, 17, 19, 31, or 37. The former are ≡ 1, 3, 7, 9 (mod 20) and the latter are ≡ 11, 13, 17, 19 (mod 20).

    Gauss's version

    The observations about −3 and 5 continue to hold: −7 is a residue modulo p if and only if p is a residue modulo 7, −11 is a residue modulo p if and only if p is a residue modulo 11, 13 is a residue (mod p) if and only if p is a residue modulo 13, etc. The more complicated-looking rules for the quadratic characters of 3 and −5, which depend upon congruences modulo 12 and 20 respectively, are simply the ones for −3 and 5 working with the first supplement.

    Example. For −5 to be a residue (mod p), either both 5 and −1 have to be residues (mod p) or they both have to be non-residues: i.e., p ≡ ±1 (mod 5) and p ≡ 1 (mod 4) or p ≡ ±2 (mod 5) and p ≡ 3 (mod 4). Using the Chinese remainder theorem these are equivalent to p ≡ 1, 9 (mod 20) or p ≡ 3, 7 (mod 20).

    The generalization of the rules for −3 and 5 is Gauss's statement of quadratic reciprocity:

    Quadratic Reciprocity (Gauss's Statement). If q 1 ( mod 4 ) then the congruence x 2 p ( mod q ) is solvable if and only if x 2 q ( mod p ) is. If q 3 ( mod 4 ) then the congruence x 2 p ( mod q ) is solvable if and only if x 2 q ( mod p ) is.

    These statements may be combined:

    Quadratic Reciprocity (Combined Statement). Define: q = ( 1 ) q 1 2 q Then the congruence x 2 p ( mod q ) is solvable if and only if x 2 q ( mod p ) is.

    Legendre's version

    Another way to organize the data is to see which primes are residues mod which other primes, as illustrated in the above table. The entry in row p column q is R if q is a quadratic residue (mod p); if it is a nonresidue the entry is N.

    If the row, or the column, or both, are ≡ 1 (mod 4) the entry is blue or green; if both row and column are ≡ 3 (mod 4), it is yellow or orange.

    The blue and green entries are symmetric around the diagonal: The entry for row p, column q is R (resp N) if and only if the entry at row q, column p, is R (resp N).

    The yellow and orange ones, on the other hand, are antisymmetric: The entry for row p, column q is R (resp N) if and only if the entry at row q, column p, is N (resp R).

    Quadratic Reciprocity (Legendre's Statement). If p or q are congruent to 1 modulo 4 then: x 2 q ( mod p ) is solvable if and only if x 2 p ( mod q ) is solvable. If p and q are congruent to 3 modulo 4 then: x 2 q ( mod p ) is solvable if and only if x 2 p ( mod q ) is not.

    It is a simple exercise to prove that Legendre's and Gauss's statements are equivalent – it requires no more than the first supplement and the facts about multiplying residues and nonresidues.

    Connection with cyclotomy

    The early proofs of quadratic reciprocity are relatively unilluminating. The situation changed when Gauss used Gauss sums to show that quadratic fields are subfields of cyclotomic fields, and implicitly deduced quadratic reciprocity from a reciprocity theorem for cyclotomic fields. His proof was cast in modern form by later algebraic number theorists. This proof served as a template for class field theory, which can be viewed as a vast generalization of quadratic reciprocity.

    Robert Langlands formulated the Langlands program, which gives a conjectural vast generalization of class field theory. He wrote:

    I confess that, as a student unaware of the history of the subject and unaware of the connection with cyclotomy, I did not find the law or its so-called elementary proofs appealing. I suppose, although I would not have (and could not have) expressed myself in this way that I saw it as little more than a mathematical curiosity, fit more for amateurs than for the attention of the serious mathematician that I then hoped to become. It was only in Hermann Weyl's book on the algebraic theory of numbers that I appreciated it as anything more.

    History and alternative statements

    There are a number of ways to state the theorem. Keep in mind that Euler and Legendre did not have Gauss's congruence notation, nor did Gauss have the Legendre symbol.

    In this article p and q always refer to distinct positive odd primes.

    Fermat

    Fermat proved (or claimed to have proved) a number of theorems about expressing a prime by a quadratic form:

    p = x 2 + y 2 p = 2  or  p 1 ( mod 4 ) p = x 2 + 2 y 2 p = 2  or  p 1 , 3 ( mod 8 ) p = x 2 + 3 y 2 p = 3  or  p 1 ( mod 3 )

    He did not state the law of quadratic reciprocity, although the cases −1, ±2, and ±3 are easy deductions from these and other of his theorems.

    He also claimed to have a proof that if the prime number p ends with 7, (in base 10) and the prime number q ends in 3, and pq ≡ 3 (mod 4), then

    p q = x 2 + 5 y 2 .

    Euler conjectured, and Lagrange proved, that

    p 1 , 9 ( mod 20 ) p = x 2 + 5 y 2 p , q 3 , 7 ( mod 20 ) p q = x 2 + 5 y 2

    Proving these and other statements of Fermat was one of the things that led mathematicians to the reciprocity theorem.

    Euler

    Translated into modern notation, Euler stated:

    1. If q ≡ 1 (mod 4) then q is a quadratic residue (mod p) if and only if pr (mod q), where r is a quadratic residue of q.
    2. If q ≡ 3 (mod 4) then q is a quadratic residue (mod p) if and only if p ≡ ±b2 (mod 4q), where b is odd and not divisible by q.

    This is equivalent to quadratic reciprocity.

    He could not prove it, but he did prove the second supplement.

    Legendre and his symbol

    Fermat proved that if p is a prime number and a is an integer,

    a p a ( mod p ) .

    Thus, if p does not divide a,

    a p 1 2 ± 1 ( mod p ) .

    Legendre lets a and A represent positive primes ≡ 1 (mod 4) and b and B positive primes ≡ 3 (mod 4), and sets out a table of eight theorems that together are equivalent to quadratic reciprocity:

    He says that since expressions of the form

    N c 1 2 ( mod c ) , gcd ( N , c ) = 1

    will come up so often he will abbreviate them as:

    ( N c ) N c 1 2 ( mod c ) = ± 1.

    This is now known as the Legendre symbol, and an equivalent definition is used today: for all integers a and all odd primes p

    ( a p ) = { 0 a 0 ( mod p ) 1 a 0 ( mod p )  and  x : a x 2 ( mod p ) 1 a 0 ( mod p )  and there is no such  x .

    Legendre's version of quadratic reciprocity

    ( p q ) = { ( q p ) p 1 ( mod 4 )  or  q 1 ( mod 4 ) ( q p )   p 3 ( mod 4 )  and  q 3 ( mod 4 )

    He notes that these can be combined:

    ( p q ) ( q p ) = ( 1 ) p 1 2 q 1 2 .

    A number of proofs, especially those based on Gauss's Lemma, explicitly calculate this formula.

    The supplementary laws using Legendre symbols

    ( 1 p ) = ( 1 ) p 1 2 = { 1 p 1 ( mod 4 ) 1 p 3 ( mod 4 ) ( 2 p ) = ( 1 ) p 2 1 8 = { 1 p 1 , 7 ( mod 8 ) 1 p 3 , 5 ( mod 8 )

    Legendre's attempt to prove reciprocity is based on a theorem of his:

    Legendre's Theorem. Let a, b and c be integers where any pair of the three are relatively prime. Moreover assume that at least one of ab, bc or ca is negative (i.e. they don't all have the same sign). If u 2 b c ( mod a ) v 2 c a ( mod b ) w 2 a b ( mod c ) are solvable then the following equation has a nontrivial solution in integers:

    Example. Theorem I is handled by letting a ≡ 1 and b ≡ 3 (mod 4) be primes and assuming that ( b a ) = 1 and, contrary the theorem, that ( a b ) = 1. Then x 2 + a y 2 b z 2 = 0 has a solution, and taking congruences (mod 4) leads to a contradiction.

    This technique doesn't work for Theorem VIII. Let bB ≡ 3 (mod 4), and assume

    ( B b ) = ( b B ) = 1.

    Then if there is another prime p ≡ 1 (mod 4) such that

    ( p b ) = ( p B ) = 1 ,

    the solvability of B x 2 + b y 2 p z 2 = 0 leads to a contradiction (mod 4). But Legendre was unable to prove there has to be such a prime p; he was later able to show that all that is required is:

    Legendre's Lemma. If a is a prime that is congruent to 1 modulo 4 then there exists a prime b such that ( a b ) = 1.

    but he couldn't prove that either. Hilbert symbol (below) discusses how techniques based on the existence of solutions to a x 2 + b y 2 + c z 2 = 0 can be made to work.

    Gauss

    Gauss first proves the supplementary laws. He sets the basis for induction by proving the theorem for ±3 and ±5. Noting that it is easier to state for −3 and +5 than it is for +3 or −5, he states the general theorem in the form:

    If p is a prime of the form 4n + 1 then p, but if p is of the form 4n + 3 then −p, is a quadratic residue (resp. nonresidue) of every prime, which, with a positive sign, is a residue (resp. nonresidue) of p. In the next sentence, he christens it the "fundamental theorem" (Gauss never used the word "reciprocity").

    Introducing the notation a R b (resp. a N b) to mean a is a quadratic residue (resp. nonresidue) (mod b), and letting a, a′, etc. represent positive primes ≡ 1 (mod 4) and b, b′, etc. positive primes ≡ 3 (mod 4), he breaks it out into the same 8 cases as Legendre:

    In the next Article he generalizes this to what are basically the rules for the Jacobi symbol (below). Letting A, A′, etc. represent any (prime or composite) positive numbers ≡ 1 (mod 4) and B, B′, etc. positive numbers ≡ 3 (mod 4):

    All of these cases take the form "if a prime is a residue (mod a composite), then the composite is a residue or nonresidue (mod the prime), depending on the congruences (mod 4)". He proves that these follow from cases 1) - 8).

    Gauss needed, and was able to prove, a lemma similar to the one Legendre needed:

    Gauss's Lemma. If p is a prime congruent to 1 modulo 8 then there exists an odd prime q such that: q < 2 p + 1 and ( p q ) = 1.

    The proof of quadratic reciprocity uses complete induction.

    Gauss's Version in Legendre Symbols. ( p q ) = { ( q p ) q 1 ( mod 4 ) ( q p ) q 3 ( mod 4 )

    These can be combined:

    Gauss's Combined Version in Legendre Symbols. Let q = ( 1 ) q 1 2 q . In other words: Then:

    A number of proofs of the theorem, especially those based on Gauss sums derive this formula. or the splitting of primes in algebraic number fields,

    Other statements

    Note that the statements in this section are equivalent to quadratic reciprocity: if, for example, Euler's version is assumed, the Legendre-Gauss version can be deduced from it, and vice versa.

    Euler's Formulation of Quadratic Reciprocity. If p ± q ( mod 4 a ) then ( a p ) = ( a q ) .

    This can be proven using Gauss's lemma.

    Quadratic Reciprocity (Gauss; Fourth Proof). Let a, b, c, ... be unequal positive odd primes, whose product is n, and let m be the number of them that are ≡ 3 (mod 4); check whether n/a is a residue of a, whether n/b is a residue of b, .... The number of nonresidues found will be even when m ≡ 0, 1 (mod 4), and it will be odd if m ≡ 2, 3 (mod 4).

    Gauss's fourth proof consists of proving this theorem (by comparing two formulas for the value of Gauss sums) and then restricting it to two primes. He then gives an example: Let a = 3, b = 5, c = 7, and d = 11. Three of these, 3, 7, and 11 ≡ 3 (mod 4), so m ≡ 3 (mod 4). 5×7×11 R 3; 3×7×11 R 5; 3×5×11 R 7;  and  3×5×7 N 11, so there are an odd number of nonresidues.

    Eisenstein's Formulation of Quadratic Reciprocity. Assume p q , p q , p p ( mod 4 ) , q q ( mod 4 ) . Then Mordell's Formulation of Quadratic Reciprocity. Let a, b and c be integers. For every prime, p, dividing abc if the congruence a x 2 + b y 2 + c z 2 0 ( mod 4 a b c p ) has a nontrivial solution, then so does:

    Jacobi symbol

    The Jacobi symbol is a generalization of the Legendre symbol; the main difference is that the bottom number has to be positive and odd, but does not have to be prime. If it is prime, the two symbols agree. It obeys the same rules of manipulation as the Legendre symbol. In particular

    ( 1 n ) = ( 1 ) n 1 2 = { 1 n 1 ( mod 4 ) 1 n 3 ( mod 4 ) ( 2 n ) = ( 1 ) n 2 1 8 = { 1 n 1 , 7 ( mod 8 ) 1 n 3 , 5 ( mod 8 )

    and if both numbers are positive and odd (this is sometimes called "Jacobi's reciprocity law"):

    ( m n ) = ( 1 ) ( m 1 ) ( n 1 ) 4 ( n m ) .

    However, if the Jacobi symbol is 1 but the denominator is not a prime, it does not necessarily follow that the numerator is a quadratic residue of the denominator. Gauss's cases 9) - 14) above can be expressed in terms of Jacobi symbols:

    ( M p ) = ( 1 ) ( p 1 ) ( M 1 ) 4 ( p M ) ,

    and since p is prime the left hand side is a Legendre symbol, and we know whether M is a residue modulo p or not.

    The formulas listed in the preceding section are true for Jacobi symbols as long as the symbols are defined. Euler's formula may be written

    ( a m ) = ( a m ± 4 a n ) , n Z , m ± 4 a n > 0.

    Example.

    ( 2 7 ) = ( 2 15 ) = ( 2 23 ) = ( 2 31 ) = = 1.

    2 is a residue modulo the primes 7, 23 and 31:

    3 2 2 ( mod 7 ) , 5 2 2 ( mod 23 ) , 8 2 2 ( mod 31 ) .

    But 2 is not a quadratic residue modulo 5, so it can't be one modulo 15. This is related to the problem Legendre had: if ( a m ) = 1 , then a is a non-residue modulo every prime in the arithmetic progression m + 4a, m + 8a, ..., if there are any primes in this series, but that wasn't proved until decades after Legendre.

    Eisenstein's formula requires relative primality conditions (which are true if the numbers are prime)

    Let a , b , a , b be positive odd integers such that: gcd ( a , b ) = gcd ( a , b ) = 1 a a ( mod 4 ) b b ( mod 4 ) Then

    Hilbert symbol

    The quadratic reciprocity law can be formulated in terms of the Hilbert symbol ( a , b ) v where a and b are any two nonzero rational numbers and v runs over all the non-trivial absolute values of the rationals (the archimedean one and the p-adic absolute values for primes p). The Hilbert symbol ( a , b ) v is 1 or −1. It is defined to be 1 if and only if the equation a x 2 + b y 2 = z 2 has a solution in the completion of the rationals at v other than x = y = z = 0 . The Hilbert reciprocity law states that ( a , b ) v , for fixed a and b and varying v, is 1 for all but finitely many v and the product of ( a , b ) v over all v is 1. (This formally resembles the residue theorem from complex analysis.)

    The proof of Hilbert reciprocity reduces to checking a few special cases, and the non-trivial cases turn out to be equivalent to the main law and the two supplementary laws of quadratic reciprocity for the Legendre symbol. There is no kind of reciprocity in the Hilbert reciprocity law; its name simply indicates the historical source of the result in quadratic reciprocity. Unlike quadratic reciprocity, which requires sign conditions (namely positivity of the primes involved) and a special treatment of the prime 2, the Hilbert reciprocity law treats all absolute values of the rationals on an equal footing. Therefore, it is a more natural way of expressing quadratic reciprocity with a view towards generalization: the Hilbert reciprocity law extends with very few changes to all global fields and this extension can rightly be considered a generalization of quadratic reciprocity to all global fields.

    Other rings

    There are also quadratic reciprocity laws in rings other than the integers.

    Gaussian integers

    In his second monograph on quartic reciprocity Gauss stated quadratic reciprocity for the ring Z [ ı ] of Gaussian integers, saying that it is a corollary of the biquadratic law in Z [ ı ] , but did not provide a proof of either theorem. Peter Gustav Lejeune Dirichlet showed that the law in Z [ ı ] can be deduced from the law for Z without using biquadratic reciprocity.

    For an odd Gaussian prime π and a Gaussian integer α relatively prime to with π , define the quadratic character for Z [ ı ] by:

    [ α π ] 2 α N π 1 2 ( mod π ) = { 1 η Z [ ı ] : α η 2 ( mod π ) 1 otherwise

    Let λ = a + b ı , μ = c + d ı be distinct Gaussian primes where a and c are odd and b and d are even. Then

    [ λ μ ] 2 = [ μ λ ] 2 , [ ı λ ] 2 = ( 1 ) b 2 , [ 1 + ı λ ] 2 = ( 2 a + b ) .

    Eisenstein integers

    Consider the following third root of unity:

    ω = 1 + 3 2 = e 2 π ı 3 .

    The ring of Eisenstein integers is Z [ ω ] . For an Eisenstein prime π , N π 3 , and an Eisenstein integer α with gcd ( α , π ) = 1 , define the quadratic character for Z [ ω ] by the formula

    [ α π ] 2 α N π 1 2 ( mod π ) = { 1 η Z [ ω ] : α η 2 ( mod π ) 1 otherwise

    Let λ = a + and μ = c + be distinct Eisenstein primes where a and c are not divisible by 3 and b and d are divisible by 3. Eisenstein proved

    [ λ μ ] 2 [ μ λ ] 2 = ( 1 ) N λ 1 2 N μ 1 2 , [ 1 ω λ ] 2 = ( a 3 ) , [ 2 λ ] 2 = ( 2 N λ ) .

    Imaginary quadratic fields

    The above laws are special cases of more general laws that hold for the ring of integers in any imaginary quadratic number field. Let k be an imaginary quadratic number field with ring of integers O k . For a prime ideal p O k with odd norm N p and α O k , define the quadratic character for O k as

    [ α p ] 2 α N p 1 2 ( mod p ) = { 1 α p  and  η O k  such that  α η 2 p 1 α p  and there is no such  η 0 α p

    for an arbitrary ideal a O k factored into prime ideals a = p 1 p n define

    [ α a ] 2 = [ α p 1 ] 2 [ α p n ] 2 ,

    and for β O k define

    [ α β ] 2 = [ α β O k ] 2 .

    Let O k = Z ω 1 Z ω 2 , i.e. { ω 1 , ω 2 } is an integral basis for O k . For ν O k with odd norm N ν , define (ordinary) integers a, b, c, d by the equations,

    ν ω 1 = a ω 1 + b ω 2 ν ω 2 = c ω 1 + d ω 2

    and a function

    χ ( ν ) := ı ( b 2 a + 2 ) c + ( a 2 b + 2 ) d + a d .

    If m = and n = are both odd, Herglotz proved

    [ μ ν ] 2 [ ν μ ] 2 = ( 1 ) m 1 2 n 1 2 χ ( μ ) m n 1 2 χ ( ν ) n m 1 2 .

    Also, if

    μ μ ( mod 4 ) and ν ν ( mod 4 )

    Then

    [ μ ν ] 2 [ ν μ ] 2 = [ μ ν ] 2 [ ν μ ] 2 .

    Polynomials over a finite field

    Let F be a finite field with q = pn elements, where p is an odd prime number and n is positive, and let F[x] be the ring of polynomials in one variable with coefficients in F. If f , g F [ x ] and f is irreducible, monic, and has positive degree, define the quadratic character for F[x] in the usual manner:

    ( g f ) = { 1 gcd ( f , g ) = 1  and  h , k F [ x ]  such that  g h 2 = k f 1 gcd ( f , g ) = 1  and  g  is not a square ( mod f ) 0 gcd ( f , g ) 1

    If f = f 1 f n is a product of monic irreducibles let

    ( g f ) = ( g f 1 ) ( g f n ) .

    Dedekind proved that if f , g F [ x ] are monic and have positive degrees,

    ( g f ) ( f g ) = ( 1 ) q 1 2 ( deg f ) ( deg g ) .

    Higher powers

    The attempt to generalize quadratic reciprocity for powers higher than the second was one of the main goals that led 19th century mathematicians, including Carl Friedrich Gauss, Peter Gustav Lejeune Dirichlet, Carl Gustav Jakob Jacobi, Gotthold Eisenstein, Richard Dedekind, Ernst Kummer, and David Hilbert to the study of general algebraic number fields and their rings of integers; specifically Kummer invented ideals in order to state and prove higher reciprocity laws.

    The ninth in the list of 23 unsolved problems which David Hilbert proposed to the Congress of Mathematicians in 1900 asked for the "Proof of the most general reciprocity law [f]or an arbitrary number field". In 1923 Emil Artin, building upon work by Philipp Furtwängler, Teiji Takagi, Helmut Hasse and others, discovered a general theorem for which all known reciprocity laws are special cases; he proved it in 1927.

    The links below provide more detailed discussions of these theorems.

    References

    Quadratic reciprocity Wikipedia