fbpx
Wikipedia

Fundamental theorem of algebra

The fundamental theorem of algebra, also called d'Alembert's theorem[1] or the d'Alembert–Gauss theorem,[2] states that every non-constant single-variable polynomial with complex coefficients has at least one complex root. This includes polynomials with real coefficients, since every real number is a complex number with its imaginary part equal to zero.

Equivalently (by definition), the theorem states that the field of complex numbers is algebraically closed.

The theorem is also stated as follows: every non-zero, single-variable, degree n polynomial with complex coefficients has, counted with multiplicity, exactly n complex roots. The equivalence of the two statements can be proven through the use of successive polynomial division.

Despite its name, there is no purely algebraic proof of the theorem, since any proof must use some form of the analytic completeness of the real numbers, which is not an algebraic concept.[3] Additionally, it is not fundamental for modern algebra; it was named when algebra was synonymous with the theory of equations.

History edit

Peter Roth [de], in his book Arithmetica Philosophica (published in 1608, at Nürnberg, by Johann Lantzenberger),[4] wrote that a polynomial equation of degree n (with real coefficients) may have n solutions. Albert Girard, in his book L'invention nouvelle en l'Algèbre (published in 1629), asserted that a polynomial equation of degree n has n solutions, but he did not state that they had to be real numbers. Furthermore, he added that his assertion holds "unless the equation is incomplete", by which he meant that no coefficient is equal to 0. However, when he explains in detail what he means, it is clear that he actually believes that his assertion is always true; for instance, he shows that the equation   although incomplete, has four solutions (counting multiplicities): 1 (twice),   and  

As will be mentioned again below, it follows from the fundamental theorem of algebra that every non-constant polynomial with real coefficients can be written as a product of polynomials with real coefficients whose degrees are either 1 or 2. However, in 1702 Leibniz erroneously said that no polynomial of the type x4 + a4 (with a real and distinct from 0) can be written in such a way. Later, Nikolaus Bernoulli made the same assertion concerning the polynomial x4 − 4x3 + 2x2 + 4x + 4, but he got a letter from Euler in 1742[5] in which it was shown that this polynomial is equal to

 

with   Also, Euler pointed out that

 

A first attempt at proving the theorem was made by d'Alembert in 1746, but his proof was incomplete. Among other problems, it assumed implicitly a theorem (now known as Puiseux's theorem), which would not be proved until more than a century later and using the fundamental theorem of algebra. Other attempts were made by Euler (1749), de Foncenex (1759), Lagrange (1772), and Laplace (1795). These last four attempts assumed implicitly Girard's assertion; to be more precise, the existence of solutions was assumed and all that remained to be proved was that their form was a + bi for some real numbers a and b. In modern terms, Euler, de Foncenex, Lagrange, and Laplace were assuming the existence of a splitting field of the polynomial p(z).

At the end of the 18th century, two new proofs were published which did not assume the existence of roots, but neither of which was complete. One of them, due to James Wood and mainly algebraic, was published in 1798 and it was totally ignored. Wood's proof had an algebraic gap.[6] The other one was published by Gauss in 1799 and it was mainly geometric, but it had a topological gap, only filled by Alexander Ostrowski in 1920, as discussed in Smale (1981).[7]

The first rigorous proof was published by Argand, an amateur mathematician, in 1806 (and revisited in 1813);[8] it was also here that, for the first time, the fundamental theorem of algebra was stated for polynomials with complex coefficients, rather than just real coefficients. Gauss produced two other proofs in 1816 and another incomplete version of his original proof in 1849.

The first textbook containing a proof of the theorem was Cauchy's Cours d'analyse de l'École Royale Polytechnique (1821). It contained Argand's proof, although Argand is not credited for it.

None of the proofs mentioned so far is constructive. It was Weierstrass who raised for the first time, in the middle of the 19th century, the problem of finding a constructive proof of the fundamental theorem of algebra. He presented his solution, which amounts in modern terms to a combination of the Durand–Kerner method with the homotopy continuation principle, in 1891. Another proof of this kind was obtained by Hellmuth Kneser in 1940 and simplified by his son Martin Kneser in 1981.

Without using countable choice, it is not possible to constructively prove the fundamental theorem of algebra for complex numbers based on the Dedekind real numbers (which are not constructively equivalent to the Cauchy real numbers without countable choice).[9] However, Fred Richman proved a reformulated version of the theorem that does work.[10]

Equivalent statements edit

There are several equivalent formulations of the theorem:

  • Every univariate polynomial of positive degree with real coefficients has at least one complex root.
  • Every univariate polynomial of positive degree with complex coefficients has at least one complex root.
    This implies immediately the previous assertion, as real numbers are also complex numbers. The converse results from the fact that one gets a polynomial with real coefficients by taking the product of a polynomial and its complex conjugate (obtained by replacing each coefficient with its complex conjugate). A root of this product is either a root of the given polynomial, or of its conjugate; in the latter case, the conjugate of this root is a root of the given polynomial.
  • Every univariate polynomial of positive degree n with complex coefficients can be factorized as
     
    where   are complex numbers.
    The n complex numbers   are the roots of the polynomial. If a root appears in several factors, it is a multiple root, and the number of its occurrences is, by definition, the multiplicity of the root.
    The proof that this statement results from the previous ones is done by recursion on n: when a root   has been found, the polynomial division by   provides a polynomial of degree   whose roots are the other roots of the given polynomial.

The next two statements are equivalent to the previous ones, although they do not involve any nonreal complex number. These statements can be proved from previous factorizations by remarking that, if r is a non-real root of a polynomial with real coefficients, its complex conjugate   is also a root, and   is a polynomial of degree two with real coefficients (this is the complex conjugate root theorem). Conversely, if one has a factor of degree two, the quadratic formula gives a root.

  • Every univariate polynomial with real coefficients of degree larger than two has a factor of degree two with real coefficients.
  • Every univariate polynomial with real coefficients of positive degree can be factored as
     
    where c is a real number and each   is a monic polynomial of degree at most two with real coefficients. Moreover, one can suppose that the factors of degree two do not have any real root.

Proofs edit

All proofs below involve some mathematical analysis, or at least the topological concept of continuity of real or complex functions. Some also use differentiable or even analytic functions. This requirement has led to the remark that the Fundamental Theorem of Algebra is neither fundamental, nor a theorem of algebra.[11]

Some proofs of the theorem only prove that any non-constant polynomial with real coefficients has some complex root. This lemma is enough to establish the general case because, given a non-constant polynomial p with complex coefficients, the polynomial

 

has only real coefficients, and, if z is a root of q, then either z or its conjugate is a root of p. Here,   is the polynomial obtained by replacing each coefficient of p with its complex conjugate; the roots of   are exactly the complex conjugates of the roots of p

Many non-algebraic proofs of the theorem use the fact (sometimes called the "growth lemma") that a polynomial function p(z) of degree n whose dominant coefficient is 1 behaves like zn when |z| is large enough. More precisely, there is some positive real number R such that

 

when |z| > R.

Real-analytic proofs edit

Even without using complex numbers, it is possible to show that a real-valued polynomial p(x): p(0) ≠ 0 of degree n > 2 can always be divided by some quadratic polynomial with real coefficients.[12] In other words, for some real-valued a and b, the coefficients of the linear remainder on dividing p(x) by x2axb simultaneously become zero.

 

where q(x) is a polynomial of degree n − 2. The coefficients Rp(x)(a, b) and Sp(x)(a, b) are independent of x and completely defined by the coefficients of p(x). In terms of representation, Rp(x)(a, b) and Sp(x)(a, b) are bivariate polynomials in a and b. In the flavor of Gauss's first (incomplete) proof of this theorem from 1799, the key is to show that for any sufficiently large negative value of b, all the roots of both Rp(x)(a, b) and Sp(x)(a, b) in the variable a are real-valued and alternating each other (interlacing property). Utilizing a Sturm-like chain that contain Rp(x)(a, b) and Sp(x)(a, b) as consecutive terms, interlacing in the variable a can be shown for all consecutive pairs in the chain whenever b has sufficiently large negative value. As Sp(a, b = 0) = p(0) has no roots, interlacing of Rp(x)(a, b) and Sp(x)(a, b) in the variable a fails at b = 0. Topological arguments can be applied on the interlacing property to show that the locus of the roots of Rp(x)(a, b) and Sp(x)(a, b) must intersect for some real-valued a and b < 0.

Complex-analytic proofs edit

Find a closed disk D of radius r centered at the origin such that |p(z)| > |p(0)| whenever |z| ≥ r. The minimum of |p(z)| on D, which must exist since D is compact, is therefore achieved at some point z0 in the interior of D, but not at any point of its boundary. The maximum modulus principle applied to 1/p(z) implies that p(z0) = 0. In other words, z0 is a zero of p(z).

A variation of this proof does not require the maximum modulus principle (in fact, a similar argument also gives a proof of the maximum modulus principle for holomorphic functions). Continuing from before the principle was invoked, if a := p(z0) ≠ 0, then, expanding p(z) in powers of zz0, we can write

 

Here, the cj are simply the coefficients of the polynomial zp(z + z0) after expansion, and k is the index of the first non-zero coefficient following the constant term. For z sufficiently close to z0 this function has behavior asymptotically similar to the simpler polynomial  . More precisely, the function

 

for some positive constant M in some neighborhood of z0. Therefore, if we define   and let   tracing a circle of radius r > 0 around z, then for any sufficiently small r (so that the bound M holds), we see that

 

When r is sufficiently close to 0 this upper bound for |p(z)| is strictly smaller than |a|, contradicting the definition of z0. Geometrically, we have found an explicit direction θ0 such that if one approaches z0 from that direction one can obtain values p(z) smaller in absolute value than |p(z0)|.

Another analytic proof can be obtained along this line of thought observing that, since |p(z)| > |p(0)| outside D, the minimum of |p(z)| on the whole complex plane is achieved at z0. If |p(z0)| > 0, then 1/p is a bounded holomorphic function in the entire complex plane since, for each complex number z, |1/p(z)| ≤ |1/p(z0)|. Applying Liouville's theorem, which states that a bounded entire function must be constant, this would imply that 1/p is constant and therefore that p is constant. This gives a contradiction, and hence p(z0) = 0.[13]

Yet another analytic proof uses the argument principle. Let R be a positive real number large enough so that every root of p(z) has absolute value smaller than R; such a number must exist because every non-constant polynomial function of degree n has at most n zeros. For each r > R, consider the number

 

where c(r) is the circle centered at 0 with radius r oriented counterclockwise; then the argument principle says that this number is the number N of zeros of p(z) in the open ball centered at 0 with radius r, which, since r > R, is the total number of zeros of p(z). On the other hand, the integral of n/z along c(r) divided by 2πi is equal to n. But the difference between the two numbers is

 

The numerator of the rational expression being integrated has degree at most n − 1 and the degree of the denominator is n + 1. Therefore, the number above tends to 0 as r → +∞. But the number is also equal to N − n and so N = n.

Another complex-analytic proof can be given by combining linear algebra with the Cauchy theorem. To establish that every complex polynomial of degree n > 0 has a zero, it suffices to show that every complex square matrix of size n > 0 has a (complex) eigenvalue.[14] The proof of the latter statement is by contradiction.

Let A be a complex square matrix of size n > 0 and let In be the unit matrix of the same size. Assume A has no eigenvalues. Consider the resolvent function

 

which is a meromorphic function on the complex plane with values in the vector space of matrices. The eigenvalues of A are precisely the poles of R(z). Since, by assumption, A has no eigenvalues, the function R(z) is an entire function and Cauchy theorem implies that

 

On the other hand, R(z) expanded as a geometric series gives:

 

This formula is valid outside the closed disc of radius   (the operator norm of A). Let   Then

 

(in which only the summand k = 0 has a nonzero integral). This is a contradiction, and so A has an eigenvalue.

Finally, Rouché's theorem gives perhaps the shortest proof of the theorem.


Topological proofs edit

 
Animation illustrating the proof on the polynomial  

Suppose the minimum of |p(z)| on the whole complex plane is achieved at z0; it was seen at the proof which uses Liouville's theorem that such a number must exist. We can write p(z) as a polynomial in z − z0: there is some natural number k and there are some complex numbers ck, ck + 1, ..., cn such that ck ≠ 0 and:

 

If p(z0) is nonzero, it follows that if a is a kth root of −p(z0)/ck and if t is positive and sufficiently small, then |p(z0 + ta)| < |p(z0)|, which is impossible, since |p(z0)| is the minimum of |p| on D.

For another topological proof by contradiction, suppose that the polynomial p(z) has no roots, and consequently is never equal to 0. Think of the polynomial as a map from the complex plane into the complex plane. It maps any circle |z| = R into a closed loop, a curve P(R). We will consider what happens to the winding number of P(R) at the extremes when R is very large and when R = 0. When R is a sufficiently large number, then the leading term zn of p(z) dominates all other terms combined; in other words,

 

When z traverses the circle   once counter-clockwise   then   winds n times counter-clockwise   around the origin (0,0), and P(R) likewise. At the other extreme, with |z| = 0, the curve P(0) is merely the single point p(0), which must be nonzero because p(z) is never zero. Thus p(0) must be distinct from the origin (0,0), which denotes 0 in the complex plane. The winding number of P(0) around the origin (0,0) is thus 0. Now changing R continuously will deform the loop continuously. At some R the winding number must change. But that can only happen if the curve P(R) includes the origin (0,0) for some R. But then for some z on that circle |z| = R we have p(z) = 0, contradicting our original assumption. Therefore, p(z) has at least one zero.

Algebraic proofs edit

These proofs of the Fundamental Theorem of Algebra must make use of the following two facts about real numbers that are not algebraic but require only a small amount of analysis (more precisely, the intermediate value theorem in both cases):

  • every polynomial with an odd degree and real coefficients has some real root;
  • every non-negative real number has a square root.

The second fact, together with the quadratic formula, implies the theorem for real quadratic polynomials. In other words, algebraic proofs of the fundamental theorem actually show that if R is any real-closed field, then its extension C = R(−1) is algebraically closed.

By induction edit

As mentioned above, it suffices to check the statement "every non-constant polynomial p(z) with real coefficients has a complex root". This statement can be proved by induction on the greatest non-negative integer k such that 2k divides the degree n of p(z). Let a be the coefficient of zn in p(z) and let F be a splitting field of p(z) over C; in other words, the field F contains C and there are elements z1, z2, ..., zn in F such that

 

If k = 0, then n is odd, and therefore p(z) has a real root. Now, suppose that n = 2km (with m odd and k > 0) and that the theorem is already proved when the degree of the polynomial has the form 2k − 1m′ with m′ odd. For a real number t, define:

 

Then the coefficients of qt(z) are symmetric polynomials in the zi with real coefficients. Therefore, they can be expressed as polynomials with real coefficients in the elementary symmetric polynomials, that is, in −a1, a2, ..., (−1)nan. So qt(z) has in fact real coefficients. Furthermore, the degree of qt(z) is n(n − 1)/2 = 2k−1m(n − 1), and m(n − 1) is an odd number. So, using the induction hypothesis, qt has at least one complex root; in other words, zi + zj + tzizj is complex for two distinct elements i and j from {1, ..., n}. Since there are more real numbers than pairs (i, j), one can find distinct real numbers t and s such that zi + zj + tzizj and zi + zj + szizj are complex (for the same i and j). So, both zi + zj and zizj are complex numbers. It is easy to check that every complex number has a complex square root, thus every complex polynomial of degree 2 has a complex root by the quadratic formula. It follows that zi and zj are complex numbers, since they are roots of the quadratic polynomial z2 −  (zi + zj)z + zizj.

Joseph Shipman showed in 2007 that the assumption that odd degree polynomials have roots is stronger than necessary; any field in which polynomials of prime degree have roots is algebraically closed (so "odd" can be replaced by "odd prime" and this holds for fields of all characteristics).[15] For axiomatization of algebraically closed fields, this is the best possible, as there are counterexamples if a single prime is excluded. However, these counterexamples rely on −1 having a square root. If we take a field where −1 has no square root, and every polynomial of degree n ∈ I has a root, where I is any fixed infinite set of odd numbers, then every polynomial f(x) of odd degree has a root (since (x2 + 1)kf(x) has a root, where k is chosen so that deg(f) + 2kI). Mohsen Aliabadi generalized[dubious ] Shipman's result in 2013, providing an independent proof that a sufficient condition for an arbitrary field (of any characteristic) to be algebraically closed is that it has a root for every polynomial of prime degree.[16]

From Galois theory edit

Another algebraic proof of the fundamental theorem can be given using Galois theory. It suffices to show that C has no proper finite field extension.[17] Let K/C be a finite extension. Since the normal closure of K over R still has a finite degree over C (or R), we may assume without loss of generality that K is a normal extension of R (hence it is a Galois extension, as every algebraic extension of a field of characteristic 0 is separable). Let G be the Galois group of this extension, and let H be a Sylow 2-subgroup of G, so that the order of H is a power of 2, and the index of H in G is odd. By the fundamental theorem of Galois theory, there exists a subextension L of K/R such that Gal(K/L) = H. As [L:R] = [G:H] is odd, and there are no nonlinear irreducible real polynomials of odd degree, we must have L = R, thus [K:R] and [K:C] are powers of 2. Assuming by way of contradiction that [K:C] > 1, we conclude that the 2-group Gal(K/C) contains a subgroup of index 2, so there exists a subextension M of C of degree 2. However, C has no extension of degree 2, because every quadratic complex polynomial has a complex root, as mentioned above. This shows that [K:C] = 1, and therefore K = C, which completes the proof.

Geometric proofs edit

There exists still another way to approach the fundamental theorem of algebra, due to J. M. Almira and A. Romero: by Riemannian geometric arguments. The main idea here is to prove that the existence of a non-constant polynomial p(z) without zeros implies the existence of a flat Riemannian metric over the sphere S2. This leads to a contradiction since the sphere is not flat.

A Riemannian surface (M, g) is said to be flat if its Gaussian curvature, which we denote by Kg, is identically null. Now, the Gauss–Bonnet theorem, when applied to the sphere S2, claims that

 

which proves that the sphere is not flat.

Let us now assume that n > 0 and

 

for each complex number z. Let us define

 

Obviously, p*(z) ≠ 0 for all z in C. Consider the polynomial f(z) = p(z)p*(z). Then f(z) ≠ 0 for each z in C. Furthermore,

 

We can use this functional equation to prove that g, given by

 

for w in C, and

 

for w ∈ S2\{0}, is a well defined Riemannian metric over the sphere S2 (which we identify with the extended complex plane C ∪ {∞}).

Now, a simple computation shows that

 

since the real part of an analytic function is harmonic. This proves that Kg = 0.

Corollaries edit

Since the fundamental theorem of algebra can be seen as the statement that the field of complex numbers is algebraically closed, it follows that any theorem concerning algebraically closed fields applies to the field of complex numbers. Here are a few more consequences of the theorem, which are either about the field of real numbers or the relationship between the field of real numbers and the field of complex numbers:

  • The field of complex numbers is the algebraic closure of the field of real numbers.
  • Every polynomial in one variable z with complex coefficients is the product of a complex constant and polynomials of the form z + a with a complex.
  • Every polynomial in one variable x with real coefficients can be uniquely written as the product of a constant, polynomials of the form x + a with a real, and polynomials of the form x2 + ax + b with a and b real and a2 − 4b < 0 (which is the same thing as saying that the polynomial x2 + ax + b has no real roots). (By the Abel–Ruffini theorem, the real numbers a and b are not necessarily expressible in terms of the coefficients of the polynomial, the basic arithmetic operations and the extraction of n-th roots.) This implies that the number of non-real complex roots is always even and remains even when counted with their multiplicity.
  • Every rational function in one variable x, with real coefficients, can be written as the sum of a polynomial function with rational functions of the form a/(x − b)n (where n is a natural number, and a and b are real numbers), and rational functions of the form (ax + b)/(x2 + cx + d)n (where n is a natural number, and a, b, c, and d are real numbers such that c2 − 4d < 0). A corollary of this is that every rational function in one variable and real coefficients has an elementary primitive.
  • Every algebraic extension of the real field is isomorphic either to the real field or to the complex field.

Bounds on the zeros of a polynomial edit

While the fundamental theorem of algebra states a general existence result, it is of some interest, both from the theoretical and from the practical point of view, to have information on the location of the zeros of a given polynomial. The simpler result in this direction is a bound on the modulus: all zeros ζ of a monic polynomial   satisfy an inequality |ζ| ≤ R, where

 

As stated, this is not yet an existence result but rather an example of what is called an a priori bound: it says that if there are solutions then they lie inside the closed disk of center the origin and radius R. However, once coupled with the fundamental theorem of algebra it says that the disk contains in fact at least one solution. More generally, a bound can be given directly in terms of any p-norm of the n-vector of coefficients   that is |ζ| ≤ Rp, where Rp is precisely the q-norm of the 2-vector   q being the conjugate exponent of p,   for any 1 ≤ p ≤ ∞. Thus, the modulus of any solution is also bounded by

 
 

for 1 < p < ∞, and in particular

 

(where we define an to mean 1, which is reasonable since 1 is indeed the n-th coefficient of our polynomial). The case of a generic polynomial of degree n,

 

is of course reduced to the case of a monic, dividing all coefficients by an ≠ 0. Also, in case that 0 is not a root, i.e. a0 ≠ 0, bounds from below on the roots ζ follow immediately as bounds from above on  , that is, the roots of

 

Finally, the distance   from the roots ζ to any point   can be estimated from below and above, seeing   as zeros of the polynomial  , whose coefficients are the Taylor expansion of P(z) at  

Let ζ be a root of the polynomial

 

in order to prove the inequality |ζ| ≤ Rp we can assume, of course, |ζ| > 1. Writing the equation as

 

and using the Hölder's inequality we find

 

Now, if p = 1, this is

 

thus

 

In the case 1 < p ≤ ∞, taking into account the summation formula for a geometric progression, we have

 

thus

 

and simplifying,

 

Therefore

 

holds, for all 1 ≤ p ≤ ∞.

See also edit

References edit

Citations edit

  1. ^ https://www.maa.org/sites/default/files/pdf/upload_library/22/Polya/07468342.di020748.02p0019l.pdf [bare URL PDF]
  2. ^ http://www.math.toronto.edu/campesat/ens/20F/14.pdf [bare URL PDF]
  3. ^ Even the proof that the equation   has a solution involves the definition of the real numbers through some form of completeness (specifically the intermediate value theorem).
  4. ^ Rare books
  5. ^ See section Le rôle d'Euler in C. Gilain's article Sur l'histoire du théorème fondamental de l'algèbre: théorie des équations et calcul intégral.
  6. ^ Concerning Wood's proof, see the article A forgotten paper on the fundamental theorem of algebra, by Frank Smithies.
  7. ^ Smale writes, "...I wish to point out what an immense gap Gauss's proof contained. It is a subtle point even today that a real algebraic plane curve cannot enter a disk without leaving. In fact, even though Gauss redid this proof 50 years later, the gap remained. It was not until 1920 that Gauss's proof was completed. In the reference Gauss, A. Ostrowski has a paper which does this and gives an excellent discussion of the problem as well..."
  8. ^ O'Connor, John J.; Robertson, Edmund F., "Jean-Robert Argand", MacTutor History of Mathematics Archive, University of St Andrews
  9. ^ For the minimum necessary to prove their equivalence, see Bridges, Schuster, and Richman; 1998; A weak countable choice principle; available from [1] 2020-02-19 at the Wayback Machine.
  10. ^ See Fred Richman; 1998; The fundamental theorem of algebra: a constructive development without choice; available from [2] 2020-02-19 at the Wayback Machine.
  11. ^ Aigner, Martin; Ziegler, Günter (2018). Proofs from the book. Springer. p. 151. ISBN 978-3-662-57264-1. OCLC 1033531310.
  12. ^ Basu, Soham (October 2021). "trictly real fundamental theorem of algebra using polynomial interlacing". Bulletin of the Australian Mathematical Society. 104 (2): 249–255. doi:10.1017/S0004972720001434. MR 4308140.
  13. ^ Ahlfors, Lars. Complex Analysis (2nd ed.). McGraw-Hill Book Company. p. 122.
  14. ^ A proof of the fact that this suffices can be seen here.
  15. ^ Shipman, J. Improving the Fundamental Theorem of Algebra. The Mathematical Intelligencer, volume 29 (2007), number 4, pp. 9–14.
  16. ^ M. Aliabadi, M. R. Darafsheh, On maximal and minimal linear matching property, Algebra and discrete mathematics, volume 15 (2013), number 2, pp. 174–178.
  17. ^ A proof of the fact that this suffices can be seen here.

Historic sources edit

  • Cauchy, Augustin-Louis (1821), Cours d'Analyse de l'École Royale Polytechnique, 1ère partie: Analyse Algébrique, Paris: Éditions Jacques Gabay (published 1992), ISBN 978-2-87647-053-8 (tr. Course on Analysis of the Royal Polytechnic Academy, part 1: Algebraic Analysis)
  • Euler, Leonhard (1751), , Histoire de l'Académie Royale des Sciences et des Belles-Lettres de Berlin, vol. 5, Berlin, pp. 222–288, archived from the original on 2008-12-24, retrieved 2008-01-28 {{citation}}: More than one of |archivedate= and |archive-date= specified (help); More than one of |archiveurl= and |archive-url= specified (help). English translation: Euler, Leonhard (1751), "Investigations on the Imaginary Roots of Equations" (PDF), Histoire de l'Académie Royale des Sciences et des Belles-Lettres de Berlin, vol. 5, Berlin, pp. 222–288
  • Gauss, Carl Friedrich (1799), Demonstratio nova theorematis omnem functionem algebraicam rationalem integram unius variabilis in factores reales primi vel secundi gradus resolvi posse, Helmstedt: C. G. Fleckeisen (tr. New proof of the theorem that every integral rational algebraic function of one variable can be resolved into real factors of the first or second degree).
  • Gauss, Carl Friedrich (1866), Carl Friedrich Gauss Werke, vol. Band III, Königlichen Gesellschaft der Wissenschaften zu Göttingen
    1. Demonstratio nova theorematis omnem functionem algebraicam rationalem integram unius variabilis in factores reales primi vel secundi gradus resolvi posse (1799), pp. 1–31., p. 1, at Google Books – first proof.
    2. Demonstratio nova altera theorematis omnem functionem algebraicam rationalem integram unius variabilis in factores reales primi vel secundi gradus resolvi posse (1815 Dec), pp. 32–56., p. 32, at Google Books – second proof.
    3. Theorematis de resolubilitate functionum algebraicarum integrarum in factores reales demonstratio tertia Supplementum commentationis praecedentis (1816 Jan), pp. 57–64., p. 57, at Google Books – third proof.
    4. Beiträge zur Theorie der algebraischen Gleichungen (1849 Juli), pp. 71–103., p. 71, at Google Books – fourth proof.
  • Kneser, Hellmuth (1940), "Der Fundamentalsatz der Algebra und der Intuitionismus", Mathematische Zeitschrift, vol. 46, pp. 287–302, doi:10.1007/BF01181442, ISSN 0025-5874, S2CID 120861330 (The Fundamental Theorem of Algebra and Intuitionism).
  • Kneser, Martin (1981), "Ergänzung zu einer Arbeit von Hellmuth Kneser über den Fundamentalsatz der Algebra", Mathematische Zeitschrift, vol. 177, no. 2, pp. 285–287, doi:10.1007/BF01214206, ISSN 0025-5874, S2CID 122310417 (tr. An extension of a work of Hellmuth Kneser on the Fundamental Theorem of Algebra).
  • Ostrowski, Alexander (1920), "Über den ersten und vierten Gaußschen Beweis des Fundamental-Satzes der Algebra", Carl Friedrich Gauss Werke Band X Abt. 2 (tr. On the first and fourth Gaussian proofs of the Fundamental Theorem of Algebra).
  • Weierstraß, Karl (1891), "Neuer Beweis des Satzes, dass jede ganze rationale Function einer Veränderlichen dargestellt werden kann als ein Product aus linearen Functionen derselben Veränderlichen", Sitzungsberichte der königlich preussischen Akademie der Wissenschaften zu Berlin, pp. 1085–1101 (tr. New proof of the theorem that every integral rational function of one variable can be represented as a product of linear functions of the same variable).

Recent literature edit

  • Almira, José María; Romero, Alfonso (2007), "Yet another application of the Gauss-Bonnet Theorem for the sphere", Bulletin of the Belgian Mathematical Society, vol. 14, pp. 341–342, MR 2341569
  • Almira, José María; Romero, Alfonso (2012), "Some Riemannian geometric proofs of the Fundamental Theorem of Algebra" (PDF), Differential Geometry – Dynamical Systems, vol. 14, pp. 1–4, MR 2914638
  • de Oliveira, Oswaldo Rio Branco (2011), "The Fundamental Theorem of Algebra: an elementary and direct proof", The Mathematical Intelligencer, vol. 33, no. 2, pp. 1–2, doi:10.1007/s00283-011-9199-2, MR 2813254, S2CID 5243991
  • de Oliveira, Oswaldo Rio Branco (2012), "The Fundamental Theorem of Algebra: from the four basic operations", The American Mathematical Monthly, vol. 119, no. 9, pp. 753–758, arXiv:1110.0165, doi:10.4169/amer.math.monthly.119.09.753, MR 2990933, S2CID 218548926
  • Fine, Benjamin; Rosenberger, Gerhard (1997), The Fundamental Theorem of Algebra, Undergraduate Texts in Mathematics, Berlin: Springer-Verlag, ISBN 978-0-387-94657-3, MR 1454356
  • Gersten, Stephen M.; Stallings, John R. (1988), "On Gauss's First Proof of the Fundamental Theorem of Algebra", Proceedings of the American Mathematical Society, vol. 103, no. 1, pp. 331–332, doi:10.1090/S0002-9939-1988-0938691-3, ISSN 0002-9939, JSTOR 2047574, MR 0938691
  • Gilain, Christian (1991), "Sur l'histoire du théorème fondamental de l'algèbre: théorie des équations et calcul intégral", Archive for History of Exact Sciences, vol. 42, no. 2, pp. 91–136, doi:10.1007/BF00496870, ISSN 0003-9519, S2CID 121468210 (tr. On the history of the fundamental theorem of algebra: theory of equations and integral calculus.)
  • Netto, Eugen; Le Vavasseur, Raymond (1916), "Les fonctions rationnelles §80–88: Le théorème fondamental", in Meyer, François; Molk, Jules (eds.), Encyclopédie des Sciences Mathématiques Pures et Appliquées, tome I, vol. 2, Éditions Jacques Gabay (published 1992), ISBN 978-2-87647-101-6 (tr. The rational functions §80–88: the fundamental theorem).
  • Remmert, Reinhold (1991), "The Fundamental Theorem of Algebra", in Ebbinghaus, Heinz-Dieter; Hermes, Hans; Hirzebruch, Friedrich (eds.), Numbers, Graduate Texts in Mathematics 123, Berlin: Springer-Verlag, ISBN 978-0-387-97497-2
  • Shipman, Joseph (2007), "Improving the Fundamental Theorem of Algebra", Mathematical Intelligencer, vol. 29, no. 4, pp. 9–14, doi:10.1007/BF02986170, ISSN 0343-6993, S2CID 123089882
  • Smale, Steve (1981), "The Fundamental Theorem of Algebra and Complexity Theory", Bulletin of the American Mathematical Society, New Series, 4 (1): 1–36, doi:10.1090/S0273-0979-1981-14858-8 [3]
  • Smith, David Eugene (1959), A Source Book in Mathematics, Dover, ISBN 978-0-486-64690-9
  • Smithies, Frank (2000), "A forgotten paper on the fundamental theorem of algebra", Notes & Records of the Royal Society, vol. 54, no. 3, pp. 333–341, doi:10.1098/rsnr.2000.0116, ISSN 0035-9149, S2CID 145593806
  • Taylor, Paul (2 June 2007), Gauss's second proof of the fundamental theorem of algebra – English translation of Gauss's second proof.
  • van der Waerden, Bartel Leendert (2003), Algebra, vol. I (7th ed.), Springer-Verlag, ISBN 978-0-387-40624-4

External links edit

fundamental, theorem, algebra, confused, with, fundamental, theorem, arithmetic, fundamental, theorem, linear, algebra, fundamental, theorem, algebra, also, called, alembert, theorem, alembert, gauss, theorem, states, that, every, constant, single, variable, p. Not to be confused with Fundamental theorem of arithmetic or Fundamental theorem of linear algebra The fundamental theorem of algebra also called d Alembert s theorem 1 or the d Alembert Gauss theorem 2 states that every non constant single variable polynomial with complex coefficients has at least one complex root This includes polynomials with real coefficients since every real number is a complex number with its imaginary part equal to zero Equivalently by definition the theorem states that the field of complex numbers is algebraically closed The theorem is also stated as follows every non zero single variable degree n polynomial with complex coefficients has counted with multiplicity exactly n complex roots The equivalence of the two statements can be proven through the use of successive polynomial division Despite its name there is no purely algebraic proof of the theorem since any proof must use some form of the analytic completeness of the real numbers which is not an algebraic concept 3 Additionally it is not fundamental for modern algebra it was named when algebra was synonymous with the theory of equations Contents 1 History 2 Equivalent statements 3 Proofs 3 1 Real analytic proofs 3 2 Complex analytic proofs 3 3 Topological proofs 3 4 Algebraic proofs 3 4 1 By induction 3 4 2 From Galois theory 3 5 Geometric proofs 4 Corollaries 5 Bounds on the zeros of a polynomial 6 See also 7 References 7 1 Citations 7 2 Historic sources 7 3 Recent literature 8 External linksHistory editPeter Roth de in his book Arithmetica Philosophica published in 1608 at Nurnberg by Johann Lantzenberger 4 wrote that a polynomial equation of degree n with real coefficients may have n solutions Albert Girard in his book L invention nouvelle en l Algebre published in 1629 asserted that a polynomial equation of degree n has n solutions but he did not state that they had to be real numbers Furthermore he added that his assertion holds unless the equation is incomplete by which he meant that no coefficient is equal to 0 However when he explains in detail what he means it is clear that he actually believes that his assertion is always true for instance he shows that the equation x 4 4 x 3 displaystyle x 4 4x 3 nbsp although incomplete has four solutions counting multiplicities 1 twice 1 i 2 displaystyle 1 i sqrt 2 nbsp and 1 i 2 displaystyle 1 i sqrt 2 nbsp As will be mentioned again below it follows from the fundamental theorem of algebra that every non constant polynomial with real coefficients can be written as a product of polynomials with real coefficients whose degrees are either 1 or 2 However in 1702 Leibniz erroneously said that no polynomial of the type x4 a4 with a real and distinct from 0 can be written in such a way Later Nikolaus Bernoulli made the same assertion concerning the polynomial x4 4x3 2x2 4x 4 but he got a letter from Euler in 1742 5 in which it was shown that this polynomial is equal to x 2 2 a x 1 7 a x 2 2 a x 1 7 a displaystyle left x 2 2 alpha x 1 sqrt 7 alpha right left x 2 2 alpha x 1 sqrt 7 alpha right nbsp with a 4 2 7 displaystyle alpha sqrt 4 2 sqrt 7 nbsp Also Euler pointed out that x 4 a 4 x 2 a 2 x a 2 x 2 a 2 x a 2 displaystyle x 4 a 4 left x 2 a sqrt 2 cdot x a 2 right left x 2 a sqrt 2 cdot x a 2 right nbsp A first attempt at proving the theorem was made by d Alembert in 1746 but his proof was incomplete Among other problems it assumed implicitly a theorem now known as Puiseux s theorem which would not be proved until more than a century later and using the fundamental theorem of algebra Other attempts were made by Euler 1749 de Foncenex 1759 Lagrange 1772 and Laplace 1795 These last four attempts assumed implicitly Girard s assertion to be more precise the existence of solutions was assumed and all that remained to be proved was that their form was a bi for some real numbers a and b In modern terms Euler de Foncenex Lagrange and Laplace were assuming the existence of a splitting field of the polynomial p z At the end of the 18th century two new proofs were published which did not assume the existence of roots but neither of which was complete One of them due to James Wood and mainly algebraic was published in 1798 and it was totally ignored Wood s proof had an algebraic gap 6 The other one was published by Gauss in 1799 and it was mainly geometric but it had a topological gap only filled by Alexander Ostrowski in 1920 as discussed in Smale 1981 7 The first rigorous proof was published by Argand an amateur mathematician in 1806 and revisited in 1813 8 it was also here that for the first time the fundamental theorem of algebra was stated for polynomials with complex coefficients rather than just real coefficients Gauss produced two other proofs in 1816 and another incomplete version of his original proof in 1849 The first textbook containing a proof of the theorem was Cauchy s Cours d analyse de l Ecole Royale Polytechnique 1821 It contained Argand s proof although Argand is not credited for it None of the proofs mentioned so far is constructive It was Weierstrass who raised for the first time in the middle of the 19th century the problem of finding a constructive proof of the fundamental theorem of algebra He presented his solution which amounts in modern terms to a combination of the Durand Kerner method with the homotopy continuation principle in 1891 Another proof of this kind was obtained by Hellmuth Kneser in 1940 and simplified by his son Martin Kneser in 1981 Without using countable choice it is not possible to constructively prove the fundamental theorem of algebra for complex numbers based on the Dedekind real numbers which are not constructively equivalent to the Cauchy real numbers without countable choice 9 However Fred Richman proved a reformulated version of the theorem that does work 10 Equivalent statements editThere are several equivalent formulations of the theorem Every univariate polynomial of positive degree with real coefficients has at least one complex root Every univariate polynomial of positive degree with complex coefficients has at least one complex root This implies immediately the previous assertion as real numbers are also complex numbers The converse results from the fact that one gets a polynomial with real coefficients by taking the product of a polynomial and its complex conjugate obtained by replacing each coefficient with its complex conjugate A root of this product is either a root of the given polynomial or of its conjugate in the latter case the conjugate of this root is a root of the given polynomial Every univariate polynomial of positive degree n with complex coefficients can be factorized as c x r 1 x r n displaystyle c x r 1 cdots x r n nbsp where c r 1 r n displaystyle c r 1 ldots r n nbsp are complex numbers The n complex numbers r 1 r n displaystyle r 1 ldots r n nbsp are the roots of the polynomial If a root appears in several factors it is a multiple root and the number of its occurrences is by definition the multiplicity of the root The proof that this statement results from the previous ones is done by recursion on n when a root r 1 displaystyle r 1 nbsp has been found the polynomial division by x r 1 displaystyle x r 1 nbsp provides a polynomial of degree n 1 displaystyle n 1 nbsp whose roots are the other roots of the given polynomial The next two statements are equivalent to the previous ones although they do not involve any nonreal complex number These statements can be proved from previous factorizations by remarking that if r is a non real root of a polynomial with real coefficients its complex conjugate r displaystyle overline r nbsp is also a root and x r x r displaystyle x r x overline r nbsp is a polynomial of degree two with real coefficients this is the complex conjugate root theorem Conversely if one has a factor of degree two the quadratic formula gives a root Every univariate polynomial with real coefficients of degree larger than two has a factor of degree two with real coefficients Every univariate polynomial with real coefficients of positive degree can be factored as c p 1 p k displaystyle cp 1 cdots p k nbsp where c is a real number and each p i displaystyle p i nbsp is a monic polynomial of degree at most two with real coefficients Moreover one can suppose that the factors of degree two do not have any real root Proofs editAll proofs below involve some mathematical analysis or at least the topological concept of continuity of real or complex functions Some also use differentiable or even analytic functions This requirement has led to the remark that the Fundamental Theorem of Algebra is neither fundamental nor a theorem of algebra 11 Some proofs of the theorem only prove that any non constant polynomial with real coefficients has some complex root This lemma is enough to establish the general case because given a non constant polynomial p with complex coefficients the polynomial q p p displaystyle q p overline p nbsp has only real coefficients and if z is a root of q then either z or its conjugate is a root of p Here p displaystyle overline p nbsp is the polynomial obtained by replacing each coefficient of p with its complex conjugate the roots of p displaystyle overline p nbsp are exactly the complex conjugates of the roots of pMany non algebraic proofs of the theorem use the fact sometimes called the growth lemma that a polynomial function p z of degree n whose dominant coefficient is 1 behaves like zn when z is large enough More precisely there is some positive real number R such that 1 2 z n lt p z lt 3 2 z n displaystyle tfrac 1 2 z n lt p z lt tfrac 3 2 z n nbsp when z gt R Real analytic proofs edit Even without using complex numbers it is possible to show that a real valued polynomial p x p 0 0 of degree n gt 2 can always be divided by some quadratic polynomial with real coefficients 12 In other words for some real valued a and b the coefficients of the linear remainder on dividing p x by x2 ax b simultaneously become zero p x x 2 a x b q x x R p x a b S p x a b displaystyle p x x 2 ax b q x x R p x a b S p x a b nbsp where q x is a polynomial of degree n 2 The coefficients Rp x a b and Sp x a b are independent of x and completely defined by the coefficients of p x In terms of representation Rp x a b and Sp x a b are bivariate polynomials in a and b In the flavor of Gauss s first incomplete proof of this theorem from 1799 the key is to show that for any sufficiently large negative value of b all the roots of both Rp x a b and Sp x a b in the variable a are real valued and alternating each other interlacing property Utilizing a Sturm like chain that contain Rp x a b and Sp x a b as consecutive terms interlacing in the variable a can be shown for all consecutive pairs in the chain whenever b has sufficiently large negative value As Sp a b 0 p 0 has no roots interlacing of Rp x a b and Sp x a b in the variable a fails at b 0 Topological arguments can be applied on the interlacing property to show that the locus of the roots of Rp x a b and Sp x a b must intersect for some real valued a and b lt 0 Complex analytic proofs edit Find a closed disk D of radius r centered at the origin such that p z gt p 0 whenever z r The minimum of p z on D which must exist since D is compact is therefore achieved at some point z0 in the interior of D but not at any point of its boundary The maximum modulus principle applied to 1 p z implies that p z0 0 In other words z0 is a zero of p z A variation of this proof does not require the maximum modulus principle in fact a similar argument also gives a proof of the maximum modulus principle for holomorphic functions Continuing from before the principle was invoked if a p z0 0 then expanding p z in powers of z z0 we can write p z a c k z z 0 k c k 1 z z 0 k 1 c n z z 0 n displaystyle p z a c k z z 0 k c k 1 z z 0 k 1 cdots c n z z 0 n nbsp Here the cj are simply the coefficients of the polynomial z p z z0 after expansion and k is the index of the first non zero coefficient following the constant term For z sufficiently close to z0 this function has behavior asymptotically similar to the simpler polynomial q z a c k z z 0 k displaystyle q z a c k z z 0 k nbsp More precisely the function p z q z z z 0 k 1 M displaystyle left frac p z q z z z 0 k 1 right leq M nbsp for some positive constant M in some neighborhood of z0 Therefore if we define 8 0 arg a p arg c k k displaystyle theta 0 arg a pi arg c k k nbsp and let z z 0 r e i 8 0 displaystyle z z 0 re i theta 0 nbsp tracing a circle of radius r gt 0 around z then for any sufficiently small r so that the bound M holds we see that p z q z r k 1 p z q z r k 1 a 1 c k r k e i arg a arg c k M r k 1 a c k r k M r k 1 displaystyle begin aligned p z amp leq q z r k 1 left frac p z q z r k 1 right 4pt amp leq left a 1 c k r k e i arg a arg c k right Mr k 1 4pt amp a c k r k Mr k 1 end aligned nbsp When r is sufficiently close to 0 this upper bound for p z is strictly smaller than a contradicting the definition of z0 Geometrically we have found an explicit direction 80 such that if one approaches z0 from that direction one can obtain values p z smaller in absolute value than p z0 Another analytic proof can be obtained along this line of thought observing that since p z gt p 0 outside D the minimum of p z on the whole complex plane is achieved at z0 If p z0 gt 0 then 1 p is a bounded holomorphic function in the entire complex plane since for each complex number z 1 p z 1 p z0 Applying Liouville s theorem which states that a bounded entire function must be constant this would imply that 1 p is constant and therefore that p is constant This gives a contradiction and hence p z0 0 13 Yet another analytic proof uses the argument principle Let R be a positive real number large enough so that every root of p z has absolute value smaller than R such a number must exist because every non constant polynomial function of degree n has at most n zeros For each r gt R consider the number 1 2 p i c r p z p z d z displaystyle frac 1 2 pi i int c r frac p z p z dz nbsp where c r is the circle centered at 0 with radius r oriented counterclockwise then the argument principle says that this number is the number N of zeros of p z in the open ball centered at 0 with radius r which since r gt R is the total number of zeros of p z On the other hand the integral of n z along c r divided by 2pi is equal to n But the difference between the two numbers is 1 2 p i c r p z p z n z d z 1 2 p i c r z p z n p z z p z d z displaystyle frac 1 2 pi i int c r left frac p z p z frac n z right dz frac 1 2 pi i int c r frac zp z np z zp z dz nbsp The numerator of the rational expression being integrated has degree at most n 1 and the degree of the denominator is n 1 Therefore the number above tends to 0 as r But the number is also equal to N n and so N n Another complex analytic proof can be given by combining linear algebra with the Cauchy theorem To establish that every complex polynomial of degree n gt 0 has a zero it suffices to show that every complex square matrix of size n gt 0 has a complex eigenvalue 14 The proof of the latter statement is by contradiction Let A be a complex square matrix of size n gt 0 and let In be the unit matrix of the same size Assume A has no eigenvalues Consider the resolvent function R z z I n A 1 displaystyle R z zI n A 1 nbsp which is a meromorphic function on the complex plane with values in the vector space of matrices The eigenvalues of A are precisely the poles of R z Since by assumption A has no eigenvalues the function R z is an entire function and Cauchy theorem implies that c r R z d z 0 displaystyle int c r R z dz 0 nbsp On the other hand R z expanded as a geometric series gives R z z 1 I n z 1 A 1 z 1 k 0 1 z k A k displaystyle R z z 1 I n z 1 A 1 z 1 sum k 0 infty frac 1 z k A k cdot nbsp This formula is valid outside the closed disc of radius A displaystyle A nbsp the operator norm of A Let r gt A displaystyle r gt A nbsp Then c r R z d z k 0 c r d z z k 1 A k 2 p i I n displaystyle int c r R z dz sum k 0 infty int c r frac dz z k 1 A k 2 pi iI n nbsp in which only the summand k 0 has a nonzero integral This is a contradiction and so A has an eigenvalue Finally Rouche s theorem gives perhaps the shortest proof of the theorem Topological proofs edit nbsp Animation illustrating the proof on the polynomial x 5 x 1 displaystyle x 5 x 1 nbsp Suppose the minimum of p z on the whole complex plane is achieved at z0 it was seen at the proof which uses Liouville s theorem that such a number must exist We can write p z as a polynomial in z z0 there is some natural number k and there are some complex numbers ck ck 1 cn such that ck 0 and p z p z 0 c k z z 0 k c k 1 z z 0 k 1 c n z z 0 n displaystyle p z p z 0 c k z z 0 k c k 1 z z 0 k 1 cdots c n z z 0 n nbsp If p z0 is nonzero it follows that if a is a kth root of p z0 ck and if t is positive and sufficiently small then p z0 ta lt p z0 which is impossible since p z0 is the minimum of p on D For another topological proof by contradiction suppose that the polynomial p z has no roots and consequently is never equal to 0 Think of the polynomial as a map from the complex plane into the complex plane It maps any circle z R into a closed loop a curve P R We will consider what happens to the winding number of P R at the extremes when R is very large and when R 0 When R is a sufficiently large number then the leading term zn of p z dominates all other terms combined in other words z n gt a n 1 z n 1 a 0 displaystyle left z n right gt left a n 1 z n 1 cdots a 0 right nbsp When z traverses the circle R e i 8 displaystyle Re i theta nbsp once counter clockwise 0 8 2 p displaystyle 0 leq theta leq 2 pi nbsp then z n R n e i n 8 displaystyle z n R n e in theta nbsp winds n times counter clockwise 0 8 2 p n displaystyle 0 leq theta leq 2 pi n nbsp around the origin 0 0 and P R likewise At the other extreme with z 0 the curve P 0 is merely the single point p 0 which must be nonzero because p z is never zero Thus p 0 must be distinct from the origin 0 0 which denotes 0 in the complex plane The winding number of P 0 around the origin 0 0 is thus 0 Now changing R continuously will deform the loop continuously At some R the winding number must change But that can only happen if the curve P R includes the origin 0 0 for some R But then for some z on that circle z R we have p z 0 contradicting our original assumption Therefore p z has at least one zero Algebraic proofs edit These proofs of the Fundamental Theorem of Algebra must make use of the following two facts about real numbers that are not algebraic but require only a small amount of analysis more precisely the intermediate value theorem in both cases every polynomial with an odd degree and real coefficients has some real root every non negative real number has a square root The second fact together with the quadratic formula implies the theorem for real quadratic polynomials In other words algebraic proofs of the fundamental theorem actually show that if R is any real closed field then its extension C R 1 is algebraically closed By induction edit As mentioned above it suffices to check the statement every non constant polynomial p z with real coefficients has a complex root This statement can be proved by induction on the greatest non negative integer k such that 2k divides the degree n of p z Let a be the coefficient of zn in p z and let F be a splitting field of p z over C in other words the field F contains C and there are elements z1 z2 zn in F such that p z a z z 1 z z 2 z z n displaystyle p z a z z 1 z z 2 cdots z z n nbsp If k 0 then n is odd and therefore p z has a real root Now suppose that n 2km with m odd and k gt 0 and that the theorem is already proved when the degree of the polynomial has the form 2k 1m with m odd For a real number t define q t z 1 i lt j n z z i z j t z i z j displaystyle q t z prod 1 leq i lt j leq n left z z i z j tz i z j right nbsp Then the coefficients of qt z are symmetric polynomials in the zi with real coefficients Therefore they can be expressed as polynomials with real coefficients in the elementary symmetric polynomials that is in a1 a2 1 nan So qt z has in fact real coefficients Furthermore the degree of qt z is n n 1 2 2k 1m n 1 and m n 1 is an odd number So using the induction hypothesis qt has at least one complex root in other words zi zj tzizj is complex for two distinct elements i and j from 1 n Since there are more real numbers than pairs i j one can find distinct real numbers t and s such that zi zj tzizj and zi zj szizj are complex for the same i and j So both zi zj and zizj are complex numbers It is easy to check that every complex number has a complex square root thus every complex polynomial of degree 2 has a complex root by the quadratic formula It follows that zi and zj are complex numbers since they are roots of the quadratic polynomial z2 zi zj z zizj Joseph Shipman showed in 2007 that the assumption that odd degree polynomials have roots is stronger than necessary any field in which polynomials of prime degree have roots is algebraically closed so odd can be replaced by odd prime and this holds for fields of all characteristics 15 For axiomatization of algebraically closed fields this is the best possible as there are counterexamples if a single prime is excluded However these counterexamples rely on 1 having a square root If we take a field where 1 has no square root and every polynomial of degree n I has a root where I is any fixed infinite set of odd numbers then every polynomial f x of odd degree has a root since x2 1 kf x has a root where k is chosen so that deg f 2k I Mohsen Aliabadi generalized dubious discuss Shipman s result in 2013 providing an independent proof that a sufficient condition for an arbitrary field of any characteristic to be algebraically closed is that it has a root for every polynomial of prime degree 16 From Galois theory edit Another algebraic proof of the fundamental theorem can be given using Galois theory It suffices to show that C has no proper finite field extension 17 Let K C be a finite extension Since the normal closure of K over R still has a finite degree over C or R we may assume without loss of generality that K is a normal extension of R hence it is a Galois extension as every algebraic extension of a field of characteristic 0 is separable Let G be the Galois group of this extension and let H be a Sylow 2 subgroup of G so that the order of H is a power of 2 and the index of H in G is odd By the fundamental theorem of Galois theory there exists a subextension L of K R such that Gal K L H As L R G H is odd and there are no nonlinear irreducible real polynomials of odd degree we must have L R thus K R and K C are powers of 2 Assuming by way of contradiction that K C gt 1 we conclude that the 2 group Gal K C contains a subgroup of index 2 so there exists a subextension M of C of degree 2 However C has no extension of degree 2 because every quadratic complex polynomial has a complex root as mentioned above This shows that K C 1 and therefore K C which completes the proof Geometric proofs edit There exists still another way to approach the fundamental theorem of algebra due to J M Almira and A Romero by Riemannian geometric arguments The main idea here is to prove that the existence of a non constant polynomial p z without zeros implies the existence of a flat Riemannian metric over the sphere S2 This leads to a contradiction since the sphere is not flat A Riemannian surface M g is said to be flat if its Gaussian curvature which we denote by Kg is identically null Now the Gauss Bonnet theorem when applied to the sphere S2 claims that S 2 K g 4 p displaystyle int mathbf S 2 K g 4 pi nbsp which proves that the sphere is not flat Let us now assume that n gt 0 and p z a 0 a 1 z a n z n 0 displaystyle p z a 0 a 1 z cdots a n z n neq 0 nbsp for each complex number z Let us define p z z n p 1 z a 0 z n a 1 z n 1 a n displaystyle p z z n p left tfrac 1 z right a 0 z n a 1 z n 1 cdots a n nbsp Obviously p z 0 for all z in C Consider the polynomial f z p z p z Then f z 0 for each z in C Furthermore f 1 w p 1 w p 1 w w 2 n p w p w w 2 n f w displaystyle f tfrac 1 w p left tfrac 1 w right p left tfrac 1 w right w 2n p w p w w 2n f w nbsp We can use this functional equation to prove that g given by g 1 f w 2 n d w 2 displaystyle g frac 1 f w frac 2 n dw 2 nbsp for w in C and g 1 f 1 w 2 n d 1 w 2 displaystyle g frac 1 left f left tfrac 1 w right right frac 2 n left d left tfrac 1 w right right 2 nbsp for w S2 0 is a well defined Riemannian metric over the sphere S2 which we identify with the extended complex plane C Now a simple computation shows that w C 1 f w 1 n K g 1 n D log f w 1 n D Re log f w 0 displaystyle forall w in mathbf C qquad frac 1 f w frac 1 n K g frac 1 n Delta log f w frac 1 n Delta text Re log f w 0 nbsp since the real part of an analytic function is harmonic This proves that Kg 0 Corollaries editSince the fundamental theorem of algebra can be seen as the statement that the field of complex numbers is algebraically closed it follows that any theorem concerning algebraically closed fields applies to the field of complex numbers Here are a few more consequences of the theorem which are either about the field of real numbers or the relationship between the field of real numbers and the field of complex numbers The field of complex numbers is the algebraic closure of the field of real numbers Every polynomial in one variable z with complex coefficients is the product of a complex constant and polynomials of the form z a with a complex Every polynomial in one variable x with real coefficients can be uniquely written as the product of a constant polynomials of the form x a with a real and polynomials of the form x2 ax b with a and b real and a2 4b lt 0 which is the same thing as saying that the polynomial x2 ax b has no real roots By the Abel Ruffini theorem the real numbers a and b are not necessarily expressible in terms of the coefficients of the polynomial the basic arithmetic operations and the extraction of n th roots This implies that the number of non real complex roots is always even and remains even when counted with their multiplicity Every rational function in one variable x with real coefficients can be written as the sum of a polynomial function with rational functions of the form a x b n where n is a natural number and a and b are real numbers and rational functions of the form ax b x2 cx d n where n is a natural number and a b c and d are real numbers such that c2 4d lt 0 A corollary of this is that every rational function in one variable and real coefficients has an elementary primitive Every algebraic extension of the real field is isomorphic either to the real field or to the complex field Bounds on the zeros of a polynomial editMain article Properties of polynomial roots While the fundamental theorem of algebra states a general existence result it is of some interest both from the theoretical and from the practical point of view to have information on the location of the zeros of a given polynomial The simpler result in this direction is a bound on the modulus all zeros z of a monic polynomial z n a n 1 z n 1 a 1 z a 0 displaystyle z n a n 1 z n 1 cdots a 1 z a 0 nbsp satisfy an inequality z R where R 1 max a 0 a n 1 displaystyle R infty 1 max a 0 ldots a n 1 nbsp As stated this is not yet an existence result but rather an example of what is called an a priori bound it says that if there are solutions then they lie inside the closed disk of center the origin and radius R However once coupled with the fundamental theorem of algebra it says that the disk contains in fact at least one solution More generally a bound can be given directly in terms of any p norm of the n vector of coefficients a a 0 a 1 a n 1 displaystyle a a 0 a 1 ldots a n 1 nbsp that is z Rp where Rp is precisely the q norm of the 2 vector 1 a p displaystyle 1 a p nbsp q being the conjugate exponent of p 1 p 1 q 1 displaystyle tfrac 1 p tfrac 1 q 1 nbsp for any 1 p Thus the modulus of any solution is also bounded by R 1 max 1 0 k lt n a k displaystyle R 1 max left 1 sum 0 leq k lt n a k right nbsp R p 1 0 k lt n a k p q p 1 q displaystyle R p left 1 left sum 0 leq k lt n a k p right frac q p right frac 1 q nbsp for 1 lt p lt and in particular R 2 0 k n a k 2 displaystyle R 2 sqrt sum 0 leq k leq n a k 2 nbsp where we define an to mean 1 which is reasonable since 1 is indeed the n th coefficient of our polynomial The case of a generic polynomial of degree n P z a n z n a n 1 z n 1 a 1 z a 0 displaystyle P z a n z n a n 1 z n 1 cdots a 1 z a 0 nbsp is of course reduced to the case of a monic dividing all coefficients by an 0 Also in case that 0 is not a root i e a0 0 bounds from below on the roots z follow immediately as bounds from above on 1 z displaystyle tfrac 1 zeta nbsp that is the roots of a 0 z n a 1 z n 1 a n 1 z a n displaystyle a 0 z n a 1 z n 1 cdots a n 1 z a n nbsp Finally the distance z z 0 displaystyle zeta zeta 0 nbsp from the roots z to any point z 0 displaystyle zeta 0 nbsp can be estimated from below and above seeing z z 0 displaystyle zeta zeta 0 nbsp as zeros of the polynomial P z z 0 displaystyle P z zeta 0 nbsp whose coefficients are the Taylor expansion of P z at z z 0 displaystyle z zeta 0 nbsp Let z be a root of the polynomial z n a n 1 z n 1 a 1 z a 0 displaystyle z n a n 1 z n 1 cdots a 1 z a 0 nbsp in order to prove the inequality z Rp we can assume of course z gt 1 Writing the equation as z n a n 1 z n 1 a 1 z a 0 displaystyle zeta n a n 1 zeta n 1 cdots a 1 zeta a 0 nbsp and using the Holder s inequality we find z n a p z n 1 z 1 q displaystyle zeta n leq a p left left zeta n 1 ldots zeta 1 right right q nbsp Now if p 1 this is z n a 1 max z n 1 z 1 a 1 z n 1 displaystyle zeta n leq a 1 max left zeta n 1 ldots zeta 1 right a 1 zeta n 1 nbsp thus z max 1 a 1 displaystyle zeta leq max 1 a 1 nbsp In the case 1 lt p taking into account the summation formula for a geometric progression we have z n a p z q n 1 z q 1 1 q a p z q n 1 z q 1 1 q a p z q n z q 1 1 q displaystyle zeta n leq a p left zeta q n 1 cdots zeta q 1 right frac 1 q a p left frac zeta qn 1 zeta q 1 right frac 1 q leq a p left frac zeta qn zeta q 1 right frac 1 q nbsp thus z n q a p q z q n z q 1 displaystyle zeta nq leq a p q frac zeta qn zeta q 1 nbsp and simplifying z q 1 a p q displaystyle zeta q leq 1 a p q nbsp Therefore z 1 a p q R p displaystyle zeta leq left left 1 a p right right q R p nbsp holds for all 1 p See also editWeierstrass factorization theorem a generalization of the theorem to other entire functions Eilenberg Niven theorem a generalization of the theorem to polynomials with quaternionic coefficients and variables Hilbert s Nullstellensatz a generalization to several variables of the assertion that complex roots exist Bezout s theorem a generalization to several variables of the assertion on the number of roots References editCitations edit https www maa org sites default files pdf upload library 22 Polya 07468342 di020748 02p0019l pdf bare URL PDF http www math toronto edu campesat ens 20F 14 pdf bare URL PDF Even the proof that the equation x 2 2 0 displaystyle x 2 2 0 nbsp has a solution involves the definition of the real numbers through some form of completeness specifically the intermediate value theorem Rare books See section Le role d Euler in C Gilain s article Sur l histoire du theoreme fondamental de l algebre theorie des equations et calcul integral Concerning Wood s proof see the article A forgotten paper on the fundamental theorem of algebra by Frank Smithies Smale writes I wish to point out what an immense gap Gauss s proof contained It is a subtle point even today that a real algebraic plane curve cannot enter a disk without leaving In fact even though Gauss redid this proof 50 years later the gap remained It was not until 1920 that Gauss s proof was completed In the reference Gauss A Ostrowski has a paper which does this and gives an excellent discussion of the problem as well O Connor John J Robertson Edmund F Jean Robert Argand MacTutor History of Mathematics Archive University of St Andrews For the minimum necessary to prove their equivalence see Bridges Schuster and Richman 1998 A weak countable choice principle available from 1 Archived 2020 02 19 at the Wayback Machine See Fred Richman 1998 The fundamental theorem of algebra a constructive development without choice available from 2 Archived 2020 02 19 at the Wayback Machine Aigner Martin Ziegler Gunter 2018 Proofs from the book Springer p 151 ISBN 978 3 662 57264 1 OCLC 1033531310 Basu Soham October 2021 trictly real fundamental theorem of algebra using polynomial interlacing Bulletin of the Australian Mathematical Society 104 2 249 255 doi 10 1017 S0004972720001434 MR 4308140 Ahlfors Lars Complex Analysis 2nd ed McGraw Hill Book Company p 122 A proof of the fact that this suffices can be seen here Shipman J Improving the Fundamental Theorem of Algebra The Mathematical Intelligencer volume 29 2007 number 4 pp 9 14 M Aliabadi M R Darafsheh On maximal and minimal linear matching property Algebra and discrete mathematics volume 15 2013 number 2 pp 174 178 A proof of the fact that this suffices can be seen here Historic sources edit Cauchy Augustin Louis 1821 Cours d Analyse de l Ecole Royale Polytechnique 1ere partie Analyse Algebrique Paris Editions Jacques Gabay published 1992 ISBN 978 2 87647 053 8 tr Course on Analysis of the Royal Polytechnic Academy part 1 Algebraic Analysis Euler Leonhard 1751 Recherches sur les racines imaginaires des equations Histoire de l Academie Royale des Sciences et des Belles Lettres de Berlin vol 5 Berlin pp 222 288 archived from the original on 2008 12 24 retrieved 2008 01 28 a href Template Citation html title Template Citation citation a More than one of archivedate and archive date specified help More than one of archiveurl and archive url specified help English translation Euler Leonhard 1751 Investigations on the Imaginary Roots of Equations PDF Histoire de l Academie Royale des Sciences et des Belles Lettres de Berlin vol 5 Berlin pp 222 288 Gauss Carl Friedrich 1799 Demonstratio nova theorematis omnem functionem algebraicam rationalem integram unius variabilis in factores reales primi vel secundi gradus resolvi posse Helmstedt C G Fleckeisen tr New proof of the theorem that every integral rational algebraic function of one variable can be resolved into real factors of the first or second degree Gauss Carl Friedrich 1866 Carl Friedrich Gauss Werke vol Band III Koniglichen Gesellschaft der Wissenschaften zu Gottingen Demonstratio nova theorematis omnem functionem algebraicam rationalem integram unius variabilis in factores reales primi vel secundi gradus resolvi posse 1799 pp 1 31 p 1 at Google Books first proof Demonstratio nova altera theorematis omnem functionem algebraicam rationalem integram unius variabilis in factores reales primi vel secundi gradus resolvi posse 1815 Dec pp 32 56 p 32 at Google Books second proof Theorematis de resolubilitate functionum algebraicarum integrarum in factores reales demonstratio tertia Supplementum commentationis praecedentis 1816 Jan pp 57 64 p 57 at Google Books third proof Beitrage zur Theorie der algebraischen Gleichungen 1849 Juli pp 71 103 p 71 at Google Books fourth proof Kneser Hellmuth 1940 Der Fundamentalsatz der Algebra und der Intuitionismus Mathematische Zeitschrift vol 46 pp 287 302 doi 10 1007 BF01181442 ISSN 0025 5874 S2CID 120861330 The Fundamental Theorem of Algebra and Intuitionism Kneser Martin 1981 Erganzung zu einer Arbeit von Hellmuth Kneser uber den Fundamentalsatz der Algebra Mathematische Zeitschrift vol 177 no 2 pp 285 287 doi 10 1007 BF01214206 ISSN 0025 5874 S2CID 122310417 tr An extension of a work of Hellmuth Kneser on the Fundamental Theorem of Algebra Ostrowski Alexander 1920 Uber den ersten und vierten Gaussschen Beweis des Fundamental Satzes der Algebra Carl Friedrich GaussWerkeBand X Abt 2 tr On the first and fourth Gaussian proofs of the Fundamental Theorem of Algebra Weierstrass Karl 1891 Neuer Beweis des Satzes dass jede ganze rationale Function einer Veranderlichen dargestellt werden kann als ein Product aus linearen Functionen derselben Veranderlichen Sitzungsberichte der koniglich preussischen Akademie der Wissenschaften zu Berlin pp 1085 1101 tr New proof of the theorem that every integral rational function of one variable can be represented as a product of linear functions of the same variable Recent literature edit Almira Jose Maria Romero Alfonso 2007 Yet another application of the Gauss Bonnet Theorem for the sphere Bulletin of the Belgian Mathematical Society vol 14 pp 341 342 MR 2341569 Almira Jose Maria Romero Alfonso 2012 Some Riemannian geometric proofs of the Fundamental Theorem of Algebra PDF Differential Geometry Dynamical Systems vol 14 pp 1 4 MR 2914638 de Oliveira Oswaldo Rio Branco 2011 The Fundamental Theorem of Algebra an elementary and direct proof The Mathematical Intelligencer vol 33 no 2 pp 1 2 doi 10 1007 s00283 011 9199 2 MR 2813254 S2CID 5243991 de Oliveira Oswaldo Rio Branco 2012 The Fundamental Theorem of Algebra from the four basic operations The American Mathematical Monthly vol 119 no 9 pp 753 758 arXiv 1110 0165 doi 10 4169 amer math monthly 119 09 753 MR 2990933 S2CID 218548926 Fine Benjamin Rosenberger Gerhard 1997 The Fundamental Theorem of Algebra Undergraduate Texts in Mathematics Berlin Springer Verlag ISBN 978 0 387 94657 3 MR 1454356 Gersten Stephen M Stallings John R 1988 On Gauss s First Proof of the Fundamental Theorem of Algebra Proceedings of the American Mathematical Society vol 103 no 1 pp 331 332 doi 10 1090 S0002 9939 1988 0938691 3 ISSN 0002 9939 JSTOR 2047574 MR 0938691 Gilain Christian 1991 Sur l histoire du theoreme fondamental de l algebre theorie des equations et calcul integral Archive for History of Exact Sciences vol 42 no 2 pp 91 136 doi 10 1007 BF00496870 ISSN 0003 9519 S2CID 121468210 tr On the history of the fundamental theorem of algebra theory of equations and integral calculus Netto Eugen Le Vavasseur Raymond 1916 Les fonctions rationnelles 80 88 Le theoreme fondamental in Meyer Francois Molk Jules eds Encyclopedie des Sciences Mathematiques Pures et Appliquees tome I vol 2 Editions Jacques Gabay published 1992 ISBN 978 2 87647 101 6 tr The rational functions 80 88 the fundamental theorem Remmert Reinhold 1991 The Fundamental Theorem of Algebra in Ebbinghaus Heinz Dieter Hermes Hans Hirzebruch Friedrich eds Numbers Graduate Texts in Mathematics 123 Berlin Springer Verlag ISBN 978 0 387 97497 2 Shipman Joseph 2007 Improving the Fundamental Theorem of Algebra Mathematical Intelligencer vol 29 no 4 pp 9 14 doi 10 1007 BF02986170 ISSN 0343 6993 S2CID 123089882 Smale Steve 1981 The Fundamental Theorem of Algebra and Complexity Theory Bulletin of the American Mathematical Society New Series 4 1 1 36 doi 10 1090 S0273 0979 1981 14858 8 3 Smith David Eugene 1959 A Source Book in Mathematics Dover ISBN 978 0 486 64690 9 Smithies Frank 2000 A forgotten paper on the fundamental theorem of algebra Notes amp Records of the Royal Society vol 54 no 3 pp 333 341 doi 10 1098 rsnr 2000 0116 ISSN 0035 9149 S2CID 145593806 Taylor Paul 2 June 2007 Gauss s second proof of the fundamental theorem of algebra English translation of Gauss s second proof van der Waerden Bartel Leendert 2003 Algebra vol I 7th ed Springer Verlag ISBN 978 0 387 40624 4External links edit nbsp Latin Wikisource has original text related to this article Gauss s first proof Algebra fundamental theorem of at Encyclopaedia of Mathematics Fundamental Theorem of Algebra a collection of proofs From the Fundamental Theorem of Algebra to Astrophysics A Harmonious Path Gauss s first proof in Latin at Google Books Gauss s first proof in Latin at Google Books Mizar system proof http mizar org version current html polynom5 html T74 Retrieved from https en wikipedia org w index php title Fundamental theorem of algebra amp oldid 1223850958, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.