Diophantine equation





Finding all right triangles with integer side-lengths is equivalent to solving the Diophantine equation a2 + b2 = c2.


In mathematics, a Diophantine equation is a polynomial equation, usually in two or more unknowns, such that only the integer solutions are sought or studied (an integer solution is a solution such that all the unknowns take integer values). A linear Diophantine equation equates the sum of two or more monomials, each of degree 1 in one of the variables, to a constant. An exponential Diophantine equation is one in which exponents on terms can be unknowns.


Diophantine problems have fewer equations than unknown variables and involve finding integers that work correctly for all equations. In more technical language, they define an algebraic curve, algebraic surface, or more general object, and ask about the lattice points on it.


The word Diophantine refers to the Hellenistic mathematician of the 3rd century, Diophantus of Alexandria, who made a study of such equations and was one of the first mathematicians to introduce symbolism into algebra. The mathematical study of Diophantine problems that Diophantus initiated is now called Diophantine analysis.


While individual equations present a kind of puzzle and have been considered throughout history, the formulation of general theories of Diophantine equations (beyond the theory of quadratic forms) was an achievement of the twentieth century.




Contents





  • 1 Examples


  • 2 Linear Diophantine equations

    • 2.1 One equation


    • 2.2 Chinese remainder theorem


    • 2.3 System of linear Diophantine equations



  • 3 Homogeneous equations

    • 3.1 Degree two

      • 3.1.1 Geometric interpretation


      • 3.1.2 Parameterization


      • 3.1.3 Example of Pythagorean triples




  • 4 Diophantine analysis

    • 4.1 Typical questions


    • 4.2 Typical problem


    • 4.3 17th and 18th centuries


    • 4.4 Hilbert's tenth problem


    • 4.5 Diophantine geometry


    • 4.6 Modern research


    • 4.7 Infinite Diophantine equations



  • 5 Exponential Diophantine equations


  • 6 See also


  • 7 Notes


  • 8 References


  • 9 Further reading


  • 10 External links




Examples


In the following Diophantine equations, w, x, y, and z are the unknowns and the other letters are given constants:














ax + by = 1This is a linear Diophantine equation.
w3 + x3 = y3 + z3The smallest nontrivial solution in positive integers is 123 + 13 = 93 + 103 = 1729. It was famously given as an evident property of 1729, a taxicab number (also named Hardy–Ramanujan number) by Ramanujan to Hardy while meeting in 1917.[1] There are infinitely many nontrivial solutions.[2]
xn + yn = znFor n = 2 there are infinitely many solutions (x,y,z): the Pythagorean triples. For larger integer values of n, Fermat's Last Theorem (initially claimed in 1637 by Fermat and proved by Wiles in 1995[3]) states there are no positive integer solutions (x,y,z).
x2ny2 = ±1This is Pell's equation, which is named after the English mathematician John Pell. It was studied by Brahmagupta in the 7th century, as well as by Fermat in the 17th century.
4/n = 1/x + 1/y + 1/zThe Erdős–Straus conjecture states that, for every positive integer n ≥ 2, there exists a solution in x, y, and z, all as positive integers. Although not usually stated in polynomial form, this example is equivalent to the polynomial equation 4xyz = yzn + xzn + xyn = n(yz + xz + xy).
x4 + y4 + z4 = w4Conjectured incorrectly by Euler to have no nontrivial solutions. Proved by Elkies to have infinitely many nontrivial solutions, with a computer search by Frye determining the smallest nontrivial solution.[4]


Linear Diophantine equations



One equation


The simplest linear Diophantine equation takes the form ax + by = c, where a, b and c are given integers. The solutions are described by the following theorem:



This Diophantine equation has a solution (where x and y are integers) if and only if c is a multiple of the greatest common divisor of a and b. Moreover, if (x, y) is a solution, then the other solutions have the form (x + kv, yku), where k is an arbitrary integer, and u and v are the quotients of a and b (respectively) by the greatest common divisor of a and b.

Proof: If d is this greatest common divisor, Bézout's identity asserts the existence of integers e and f such that ae + bf = d. If c is a multiple of d, then c = dh for some integer h, and (eh, fh) is a solution. On the other hand, for every pair of integers x and y, the greatest common divisor d of a and b divides ax + by. Thus, if the equation has a solution, then c must be a multiple of d. If a = ud and b = vd, then for every solution (x, y), we have



a(x + kv) + b(yku) = ax + by + k(avbu) = ax + by + k(udvvdu) = ax + by,

showing that (x + kv, yku) is another solution. Finally, given two solutions such that ax1 + by1 = ax2 + by2 = c, one deduces that u(x2x1) + v(y2y1) = 0. As u and v are coprime, Euclid's lemma shows that v divides x2x1, and thus that
there exists an integer k such that x2x1 = kv and y2y1 = −ku. Therefore, x2 = x1 + kv and y2 = y1ku, which completes the proof.



Chinese remainder theorem


The Chinese remainder theorem describes an important class of linear Diophantine systems of equations: let n1, …, nk be k pairwise coprime integers greater than one, a1, …, ak be k arbitrary integers, and N be the product n1 ··· nk. The Chinese remainder theorem asserts that the following linear Diophantine system has exactly one solution (x, x1, …, xk) such that 0 ≤ x < N, and that the other solutions are obtained by adding to x a multiple of N:


x=a1+n1x1⋮x=ak+nkxkdisplaystyle beginalignedx&=a_1+n_1,x_1\&vdots \x&=a_k+n_k,x_kendalignedbeginalignedx&=a_1+n_1,x_1\&vdots \x&=a_k+n_k,x_kendaligned


System of linear Diophantine equations


More generally, every system of linear Diophantine equations may be solved by computing the Smith normal form of its matrix, in a way that is similar to the use of the reduced row echelon form to solve a system of linear equations over a field. Using matrix notation every system of linear Diophantine equations may be written



AX = C,

where A is an m × n matrix of integers, X is an n × 1 column matrix of unknowns and C is an m × 1 column matrix of integers.


The computation of the Smith normal form of A provides two unimodular matrices (that is matrices that are invertible over the integers and have ±1 as determinant) U and V of respective dimensions m × m and n × n, such that the matrix


B = [bi,j] = UAV

is such that bi,i is not zero for i not greater than some integer k, and all the other entries are zero. The system to be solved may thus be rewritten as



B (V−1X) = UC.

Calling yi the entries of V−1X and di those of D = UC, this leads to the system



bi,iyi = di for 1 ≤ ik,


0 yi = di for k < in.

This system is equivalent to the given one in the following sense: A column matrix of integers x is a solution of the given system if and only if x = Vy for some column matrix of integers y such that By = D.


It follows that the system has a solution if and only if bi,i divides di for ik and di = 0 for i > k. If this condition is fulfilled, the solutions of the given system are


V[d1b1,1⋮dkbk,khk+1⋮hn],displaystyle V,left[beginarraycfrac d_1b_1,1\vdots \frac d_kb_k,k\h_k+1\vdots \h_nendarrayright],,V,left[beginarraycfrac d_1b_1,1\vdots \frac d_kb_k,k\h_k+1\vdots \h_nendarrayright],,

where hk+1, ..., hn are arbitrary integers.


Hermite normal form may also be used for solving systems of linear Diophantine equations. However, Hermite normal form does not directly provide the solutions; to get the solutions from the Hermite normal form, one has to successively solve several linear equations. Nevertheless, Richard Zippel wrote that the Smith normal form "is somewhat more than is actually needed to solve linear diophantine equations. Instead of reducing the equation to diagonal form, we only need to make it triangular, which is called the Hermite normal form. The Hermite normal form is substantially easier to compute than the Smith normal form."[5]


Integer linear programming amounts to finding some integer solutions (optimal in some sense) of linear systems that include also inequations. Thus systems of linear Diophantine equations are basic in this context, and textbooks on integer programming usually have a treatment of systems of linear Diophantine equations.[6]



Homogeneous equations


A homogeneous Diophantine equation is a Diophantine equation that is defined by a homogeneous polynomial. A typical such equation is the equation of Fermat's Last Theorem


xd+yd−zd=0.displaystyle x^d+y^d-z^d=0.displaystyle x^d+y^d-z^d=0.

As a homogeneous polynomial in n indeterminates defines a hypersurface in the projective space of dimension n – 1, solving a homogeneous Diophantine equation is the same as finding the rational points of a projective hypersurface.


Solving a homogeneous Diophantine equation is generally a very difficult problem, even in the simplest non-trivial case of three indeterminates (in the case of two indeterminates the problem is equivalent with testing if a rational number is the dth power of another rational number). A witness of the difficulty of the problem is Fermat's Last Theorem (for d > 2, there is no integer solution of the above equation), which needed more than three centuries of mathematicians efforts for being solved.


For degrees higher than three, most known results are theorems asserting that there are no solutions (for example Fermat's Last Theorem) or that the number of solutions is finite (for example Falting's theorem).


For the degree three, there are general solving methods, which work on almost all equations that are encountered in practice, but no algorithm is known that works for every cubic equation.[citation needed]



Degree two


Homogeneous Diophantine equations of degree two are easier to solve. The standard solving method proceed in two steps. One has first to find one solution, or to prove that there is no solution. When a solution has been found, all solutions are then deduced.


For proving that there is no solution, one may reduce the equation modulo p. For example, the Diophantine equation


x2+y2=3z2,displaystyle x^2+y^2=3z^2,displaystyle x^2+y^2=3z^2,

does not have any other solution than the trivial solution (0, 0, 0). In fact, by dividing x, y and z by their greatest common divisor, one may suppose that they are coprime. The squares modulo 4 are congruent to 0 and 1. Thus the left-hand side of the equation is congruent to 0, 1, or 2, and the right-hand side is congruent to 0 or 3. Thus the equality may be obtained only if x, y and z are all even, and are thus not coprime. Thus the only solution is the trivial solution (0, 0, 0). This shows that there is no rational point on a circle of radius 3,displaystyle sqrt 3,displaystyle sqrt 3, centered at the origin.


More generally, Hasse principle allows deciding whether a homogeneous Diophantine equation of degree two has an integer solution, and computing a solution if there exist.


If a non-trivial integer solution is known, one may produce all other solutions in the following way.



Geometric interpretation


Let


Q(x1,…,xn)=0displaystyle Q(x_1,ldots ,x_n)=0displaystyle Q(x_1,ldots ,x_n)=0

be a homogeneous Diophantine equation, where Q(x1,…,xn)displaystyle Q(x_1,ldots ,x_n)displaystyle Q(x_1,ldots ,x_n) is a quadratic form (that is, a homogeneous polynomial of degree 2), with integer coefficients. The trivial solution is the solution where all xidisplaystyle x_ix_i are zero. If (a1,…,an)displaystyle (a_1,ldots ,a_n)displaystyle (a_1,ldots ,a_n) is a non-trivial integer solution of this equation, then (a1,…,an)displaystyle left(a_1,ldots ,a_nright)displaystyle left(a_1,ldots ,a_nright) are the homogeneous coordinates of a rational point of the hypersurface defined by Q. Conversely, if (p1q,…,pnq)displaystyle left(frac p_1q,ldots ,frac p_nqright)displaystyle left(frac p_1q,ldots ,frac p_nqright) are homogeneous coordinates of a rational point of this hypersurface, where q,p1,…,pndisplaystyle q,p_1,ldots ,p_ndisplaystyle q,p_1,ldots ,p_n are integers, then (p1,…,pn)displaystyle left(p_1,ldots ,p_nright)displaystyle left(p_1,ldots ,p_nright) is an integer solution of the Diophantine equation. Moreover, the integer solutions that define a given rational point are all sequences of the form


(kp1d,…,kpnd),displaystyle left(kfrac p_1d,ldots ,kfrac p_ndright),displaystyle left(kfrac p_1d,ldots ,kfrac p_ndright),

where k is any integer, and d is the greatest common divisor of the p1.displaystyle p_1.displaystyle p_1.


It follows that solving the Diophantine equation Q(x1,…,xn)=0displaystyle Q(x_1,ldots ,x_n)=0displaystyle Q(x_1,ldots ,x_n)=0 is completely reduced to finding the rational points of the corresponding projective hypersurface.



Parameterization


Let now A=(a1,…,an)displaystyle A=left(a_1,ldots ,a_nright)displaystyle A=left(a_1,ldots ,a_nright) be an integer solution of the equation Q(x1,…,xn)=0.displaystyle Q(x_1,ldots ,x_n)=0.displaystyle Q(x_1,ldots ,x_n)=0. As Q is a polynomial of degree two, a line passing through A crosses the hypersurface at a single other point, which is rational if and only if the line is rational (that is, if the line is defined by rational parameters). This allows parameterizing the hypersurface by the lines passing through A, and the rational points are the those that are obtained from rational lines, that is, those that correspond to rational values of the parameters.


More precisely, one may proceed as follows.


By permuting the indices, one may suppose, without loss of generality that an≠0.displaystyle a_nneq 0.a_nne 0. Then one may pass to the affine case by considering the affine hypersurface defined by


q(x1,…,xn−1)=Q(x1,…,xn−1,1),displaystyle q(x_1,ldots ,x_n-1)=Q(x_1,ldots ,x_n-1,1),displaystyle q(x_1,ldots ,x_n-1)=Q(x_1,ldots ,x_n-1,1),

which has the rational point


R=(r1,…,rn−1)=(a1an,…,an−1an).displaystyle R=(r_1,ldots ,r_n-1)=left(frac a_1a_n,ldots ,frac a_n-1a_nright).displaystyle R=(r_1,ldots ,r_n-1)=left(frac a_1a_n,ldots ,frac a_n-1a_nright).

If this rational point is a singular point, that is if all partial derivatives are zero at R, all line passing through R are contained in the hypersurface, and one has a cone. The change of variables


yi=xi−ridisplaystyle y_i=x_i-r_idisplaystyle y_i=x_i-r_i

does not change the rational points, and transforms q into a homogeneous polynomial in n – 1 variables. In this case, the problem may thus be solved by applying the method to an equation with fewer variables.


If the polynomial q is a product of linear polynomials (possibly with non-rational coefficients), then it defines two hyperplanes. The intersection of these hyperplanes is a rational flat, and contains rational singular points. This case is thus a special instance of the preceding case.


In the general case, let consider the parametric equation of a line passing through R:


x2=r2+t2(x1−r1)⋮xn−1=rn−1+tn−1(x1−r1).displaystyle beginalignedx_2&=r_2+t_2(x_1-r_1)\vdots &\x_n-1&=r_n-1+t_n-1(x_1-r_1).endaligneddisplaystyle beginalignedx_2&=r_2+t_2(x_1-r_1)\vdots &\x_n-1&=r_n-1+t_n-1(x_1-r_1).endaligned

Substituting this in q, one gets a polynomial of degree two in x1,displaystyle x_1,displaystyle x_1, that is zero for x1=r1.displaystyle x_1=r_1.displaystyle x_1=r_1. It is thus divisible by x1−r1,displaystyle x_1-r_1,displaystyle x_1-r_1,. The quotient is linear in x1,displaystyle x_1,displaystyle x_1, and may be solved for expressing x1displaystyle x_1x_1 as a quotient of two polynomials of degree at most two in t2,…,tn−1,displaystyle t_2,ldots ,t_n-1,displaystyle t_2,ldots ,t_n-1, with integer coefficients:


x1=f1(t2,…,tn−1)fn(t2,…,tn−1).displaystyle x_1=frac f_1(t_2,ldots ,t_n-1)f_n(t_2,ldots ,t_n-1).displaystyle x_1=frac f_1(t_2,ldots ,t_n-1)f_n(t_2,ldots ,t_n-1).

Substituting this in the expressions for x2,…,xn−1,displaystyle x_2,ldots ,x_n-1,displaystyle x_2,ldots ,x_n-1, one gets, for i = 1, ..., n – 1,


xi=fi(t2,…,tn−1)fn(t2,…,tn−1),displaystyle x_i=frac f_i(t_2,ldots ,t_n-1)f_n(t_2,ldots ,t_n-1),displaystyle x_i=frac f_i(t_2,ldots ,t_n-1)f_n(t_2,ldots ,t_n-1),

where f1,…,fndisplaystyle f_1,ldots ,f_ndisplaystyle f_1,ldots ,f_n are polynomials of degree at most two with integer coefficients.


Then, one can return to the homogeneous case. Let, for i = 1, ..., n,


Fi(t1,…,tn−1)=t12fi(t2t1,…,tn−1t1),displaystyle F_i(t_1,ldots ,t_n-1)=t_1^2f_ileft(frac t_2t_1,ldots ,frac t_n-1t_1right),displaystyle F_i(t_1,ldots ,t_n-1)=t_1^2f_ileft(frac t_2t_1,ldots ,frac t_n-1t_1right),

be the homogenization of fi.displaystyle f_i.displaystyle f_i. These quadratic polynomials with integer coefficients form a parameterization of the projective hypersurface defined by Q:


x1=F1(t1,…,tn−1)⋮xn=Fn(t1,…,tn−1).displaystyle beginalignedx_1&=F_1(t_1,ldots ,t_n-1)\vdots &\x_n&=F_n(t_1,ldots ,t_n-1).endaligneddisplaystyle beginalignedx_1&=F_1(t_1,ldots ,t_n-1)\vdots &\x_n&=F_n(t_1,ldots ,t_n-1).endaligned

A point of the projective hypersurface defined by Q is rational if and only if it may be obtained from rational values of t1,…,tn−1.displaystyle t_1,ldots ,t_n-1.displaystyle t_1,ldots ,t_n-1. As F1,…,Fndisplaystyle F_1,ldots ,F_ndisplaystyle F_1,ldots ,F_n are homogeneous polynomials, the point is not changed if all tidisplaystyle t_it_i are multiplied by the same rational number. Thus, one may suppose that t1,…,tn−1displaystyle t_1,ldots ,t_n-1displaystyle t_1,ldots ,t_n-1 are coprime integers. It follows that the integer solutions of the Diophantine equation are exactly the sequences (x1,…,xn)displaystyle (x_1,ldots ,x_n)(x_1, ldots, x_n) where, for i = 1, ..., n,


xi=kFi(t1,…,tn−1)d,displaystyle x_i=k,frac F_i(t_1,ldots ,t_n-1)d,displaystyle x_i=k,frac F_i(t_1,ldots ,t_n-1)d,

where k is an integer, t1,…,tn−1displaystyle t_1,ldots ,t_n-1displaystyle t_1,ldots ,t_n-1 are coprime integers, and d is the greatest common divisor of the n integers Fi(t1,…,tn−1).displaystyle F_i(t_1,ldots ,t_n-1).displaystyle F_i(t_1,ldots ,t_n-1).


One could hope that the coprimality of the tidisplaystyle t_it_i could imply that d = 1. Unfortunately this is not the case, as shown in the next section.



Example of Pythagorean triples


The equation


x2+y2−z2=0displaystyle x^2+y^2-z^2=0x^2+y^2-z^2=0

is probably the first homogeneous Diophantine equation of degree two that has been studied. Its solutions are the Pythagorean triples. This is also the homogeneous equation of the unit circle. In this section, we show how the above method allows retrieving Euclid's formula for generating Pythagorean triples.


For retrieving exactly Euclid's formula, we start from the solution (-1, 0, 1), corresponding to the point (-1, 0) of the unit circle. A line passing through this point may be parameterized by its slope:


y=t(x+1).displaystyle y=t(x+1).displaystyle y=t(x+1).

Putting this in the circle equation


x2+y2−1=0,displaystyle x^2+y^2-1=0,displaystyle x^2+y^2-1=0,

one gets


x2−1+t2(x+1)2=0.displaystyle x^2-1+t^2(x+1)^2=0.displaystyle x^2-1+t^2(x+1)^2=0.

Dividing by x + 1, results in


x−1+t2(x+1)=0,displaystyle x-1+t^2(x+1)=0,displaystyle x-1+t^2(x+1)=0,

which is easy to solve in x:


x=1−t21+t2.displaystyle x=frac 1-t^21+t^2.displaystyle x=frac 1-t^21+t^2.

It follows


y=t(x+1)=2t1+t2.displaystyle y=t(x+1)=frac 2t1+t^2.displaystyle y=t(x+1)=frac 2t1+t^2.

Homogenizing as described above one gets all solutions as


x=ks2−t2dy=k2stdz=ks2+t2d,displaystyle beginalignedx&=k,frac s^2-t^2d\y&=k,frac 2std\z&=k,frac s^2+t^2d,endaligneddisplaystyle beginalignedx&=k,frac s^2-t^2d\y&=k,frac 2std\z&=k,frac s^2+t^2d,endaligned

where k is any integer, s and t are coprime integers, and d is the greatest common divisor of the three numerators. In fact, d = 2 if s and t are both odd, and d = 1 if one is odd and the other is even.


The primitive triples are the solutions where k = 1 and s > t > 0.


This description of the solutions differs slightly from Euclid's formula because Euclid's formula considers only the solutions such that x, y and z are all positive, and does not distinguish between two triples that differ by the exchange of x and y,



Diophantine analysis



Typical questions


The questions asked in Diophantine analysis include:


  1. Are there any solutions?

  2. Are there any solutions beyond some that are easily found by inspection?

  3. Are there finitely or infinitely many solutions?

  4. Can all solutions be found in theory?

  5. Can one in practice compute a full list of solutions?

These traditional problems often lay unsolved for centuries, and mathematicians gradually came to understand their depth (in some cases), rather than treat them as puzzles.



Typical problem


The given information is that a father's age is 1 less than twice that of his son, and that the digits AB making up the father's age are reversed in the son's age (i.e. BA). This leads to the equation 10A + B = 2(10B + A) − 1, thus 19B − 8A = 1. Inspection gives the result A = 7, B = 3, and thus AB equals 73 years and BA equals 37 years. One may easily show that there is not any other solution with A and B positive integers less than 10.


Many well known puzzles in the field of recreational mathematics lead to diophantine equations. Examples include the Cannonball problem, Archimedes's cattle problem and The monkey and the coconuts.



17th and 18th centuries


In 1637, Pierre de Fermat scribbled on the margin of his copy of Arithmetica: "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." Stated in more modern language, "The equation an + bn = cn has no solutions for any n higher than 2." And then he wrote, intriguingly: "I have discovered a truly marvelous proof of this proposition, which this margin is too narrow to contain." Such a proof eluded mathematicians for centuries, however, and as such his statement became famous as Fermat's Last Theorem. It wasn't until 1995 that it was proven by the British mathematician Andrew Wiles.


In 1657, Fermat attempted to solve the Diophantine equation 61x2 + 1 = y2 (solved by Brahmagupta over 1000 years earlier). The equation was eventually solved by Euler in the early 18th century, who also solved a number of other Diophantine equations. The smallest solution of this equation in positive integers is x = 226153980, y = 1766319049 (see Chakravala method).



Hilbert's tenth problem



In 1900, David Hilbert proposed the solvability of all Diophantine equations as the tenth of his fundamental problems. In 1970, Yuri Matiyasevich solved it negatively, by proving that a general algorithm for solving all Diophantine equations cannot exist.



Diophantine geometry


Diophantine geometry, which is the application of techniques from algebraic geometry in this field, has continued to grow as a result; since treating arbitrary equations is a dead end, attention turns to equations that also have a geometric meaning. The central idea of Diophantine geometry is that of a rational point, namely a solution to a polynomial equation or a system of polynomial equations, which is a vector in a prescribed field K, when K is not algebraically closed.



Modern research


One of the few general approaches is through the Hasse principle. Infinite descent is the traditional method, and has been pushed a long way.


The depth of the study of general Diophantine equations is shown by the characterisation of Diophantine sets as equivalently described as recursively enumerable. In other words, the general problem of Diophantine analysis is blessed or cursed with universality, and in any case is not something that will be solved except by re-expressing it in other terms.


The field of Diophantine approximation deals with the cases of Diophantine inequalities. Here variables are still supposed to be integral, but some coefficients may be irrational numbers, and the equality sign is replaced by upper and lower bounds.


The most celebrated single question in the field, the conjecture known as Fermat's Last Theorem, was solved by Andrew Wiles[7] but using tools from algebraic geometry developed during the last century rather than within number theory where the conjecture was originally formulated. Other major results, such as Faltings's theorem, have disposed of old conjectures.



Infinite Diophantine equations


An example of an infinite diophantine equation is:



n = a2 + 2b2 + 3c2 + 4d2 + 5e2 + …,

which can be expressed as "How many ways can a given integer n be written as the sum of a square plus twice a square plus thrice a square and so on?" The number of ways this can be done for each n forms an integer sequence. Infinite Diophantine equations are related to theta functions and infinite dimensional lattices. This equation always has a solution for any positive n. Compare this to:



n = a2 + 4b2 + 9c2 + 16d2 + 25e2 + …,

which does not always have a solution for positive n.



Exponential Diophantine equations


If a Diophantine equation has as an additional variable or variables occurring as exponents, it is an exponential Diophantine equation. Examples include the Ramanujan–Nagell equation, 2n − 7 = x2, and the equation of the Fermat-Catalan conjecture and Beal's conjecture, am + bn = ck with inequality restrictions on the exponents. A general theory for such equations is not available; particular cases such as Catalan's conjecture have been tackled. However, the majority are solved via ad hoc methods such as Størmer's theorem or even trial and error.



See also



  • Kuṭṭaka, Aryabhata's algorithm for solving linear Diophantine equations in two unknowns


Notes




  1. ^ "Quotations by Hardy". Gap.dcs.st-and.ac.uk. Retrieved 20 November 2012..mw-parser-output cite.citationfont-style:inherit.mw-parser-output qquotes:"""""""'""'".mw-parser-output code.cs1-codecolor:inherit;background:inherit;border:inherit;padding:inherit.mw-parser-output .cs1-lock-free abackground:url("//upload.wikimedia.org/wikipedia/commons/thumb/6/65/Lock-green.svg/9px-Lock-green.svg.png")no-repeat;background-position:right .1em center.mw-parser-output .cs1-lock-limited a,.mw-parser-output .cs1-lock-registration abackground:url("//upload.wikimedia.org/wikipedia/commons/thumb/d/d6/Lock-gray-alt-2.svg/9px-Lock-gray-alt-2.svg.png")no-repeat;background-position:right .1em center.mw-parser-output .cs1-lock-subscription abackground:url("//upload.wikimedia.org/wikipedia/commons/thumb/a/aa/Lock-red-alt-2.svg/9px-Lock-red-alt-2.svg.png")no-repeat;background-position:right .1em center.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registrationcolor:#555.mw-parser-output .cs1-subscription span,.mw-parser-output .cs1-registration spanborder-bottom:1px dotted;cursor:help.mw-parser-output .cs1-hidden-errordisplay:none;font-size:100%.mw-parser-output .cs1-visible-errorfont-size:100%.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration,.mw-parser-output .cs1-formatfont-size:95%.mw-parser-output .cs1-kern-left,.mw-parser-output .cs1-kern-wl-leftpadding-left:0.2em.mw-parser-output .cs1-kern-right,.mw-parser-output .cs1-kern-wl-rightpadding-right:0.2em


  2. ^ Everest, G.; Ward, Thomas (2006), An Introduction to Number Theory, Graduate Texts in Mathematics, 232, Springer, p. 117, ISBN 9781846280443.


  3. ^ Wiles, Andrew (1995). "Modular elliptic curves and Fermat's Last Theorem" (PDF). Annals of Mathematics. Annals of Mathematics. 141 (3): 443–551. doi:10.2307/2118559. JSTOR 2118559. OCLC 37032255.


  4. ^ Noam Elkies (1988). "On A4 + B4 + C4 = D4". Mathematics of Computation. 51 (184): 825–835. doi:10.2307/2008781. JSTOR 2008781. MR 0930224.


  5. ^ Richard Zippel (1993). Effective Polynomial Computation. Springer Science & Business Media. p. 50. ISBN 978-0-7923-9375-7.


  6. ^ Alexander Bockmayr, Volker Weispfenning (2001). "Solving Numerical Constraints". In John Alan Robinson and Andrei Voronkov. Handbook of Automated Reasoning Volume I. Elsevier and MIT Press. p. 779.
    ISBN 0-444-82949-0 (Elsevier)
    ISBN 0-262-18221-1 (MIT Press).



  7. ^ Solving Fermat: Andrew Wiles




References



  • Mordell, L. J. (1969). Diophantine equations. Pure and Applied Mathematics. 30. Academic Press. ISBN 0-12-506250-8. Zbl 0188.34503.


  • Schmidt, Wolfgang M. (1991). Diophantine approximations and Diophantine equations. Lecture Notes in Mathematics. 1467. Berlin: Springer-Verlag. ISBN 3-540-54058-X. Zbl 0754.11020.


  • Shorey, T. N.; Tijdeman, R. (1986). Exponential Diophantine equations. Cambridge Tracts in Mathematics. 87. Cambridge University Press. ISBN 0-521-26826-5. Zbl 0606.10011.


  • Smart, Nigel P. (1998). The algorithmic resolution of Diophantine equations. London Mathematical Society Student Texts. 41. Cambridge University Press. ISBN 0-521-64156-X. Zbl 0907.11001.


  • Stillwell, John (2004). Mathematics and its History (Second ed.). Springer Science + Business Media Inc. ISBN 0-387-95336-1.


Further reading



  • Dickson, Leonard Eugene (2005) [1920]. History of the Theory of Numbers. Volume II: Diophantine analysis. Mineola, NY: Dover Publications. ISBN 978-0-486-44233-4. MR 0245500. Zbl 1214.11002.


External links



  • Diophantine Equation. From MathWorld at Wolfram Research.


  • Diophantine Equation. From PlanetMath.


  • Hazewinkel, Michiel, ed. (2001) [1994], "Diophantine equations", Encyclopedia of Mathematics, Springer Science+Business Media B.V. / Kluwer Academic Publishers, ISBN 978-1-55608-010-4


  • Dario Alpern's Online Calculator. Retrieved 18 March 2009








這個網誌中的熱門文章

Barbados

How to read a connectionString WITH PROVIDER in .NET Core?

Node.js Script on GitHub Pages or Amazon S3