Lindemann–Weierstrass theorem
   HOME

TheInfoList



OR:

In transcendental number theory, the Lindemann–Weierstrass theorem is a result that is very useful in establishing the
transcendence Transcendence, transcendent, or transcendental may refer to: Mathematics * Transcendental number, a number that is not the root of any polynomial with rational coefficients * Algebraic element or transcendental element, an element of a field exten ...
of numbers. It states the following: In other words, the extension field \mathbb(e^, \dots, e^) has
transcendence degree In abstract algebra, the transcendence degree of a field extension ''L'' / ''K'' is a certain rather coarse measure of the "size" of the extension. Specifically, it is defined as the largest cardinality of an algebraically independent subset of ...
over \mathbb. An equivalent formulation , is the following: This equivalence transforms a linear relation over the algebraic numbers into an algebraic relation over \mathbb by using the fact that a symmetric polynomial whose arguments are all conjugates of one another gives a rational number. The theorem is named for Ferdinand von Lindemann and
Karl Weierstrass Karl Theodor Wilhelm Weierstrass (german: link=no, Weierstraß ; 31 October 1815 – 19 February 1897) was a German mathematician often cited as the "father of modern analysis". Despite leaving university without a degree, he studied mathematics ...
. Lindemann proved in 1882 that is transcendental for every non-zero algebraic number thereby establishing that is transcendental (see below). Weierstrass proved the above more general statement in 1885. The theorem, along with the
Gelfond–Schneider theorem In mathematics, the Gelfond–Schneider theorem establishes the transcendence of a large class of numbers. History It was originally proved independently in 1934 by Aleksandr Gelfond and Theodor Schneider. Statement : If ''a'' and ''b'' a ...
, is extended by
Baker's theorem In transcendental number theory, a mathematical discipline, Baker's theorem gives a lower bound for the absolute value of linear combinations of logarithms of algebraic numbers. The result, proved by , subsumed many earlier results in transcendent ...
, and all of these would be further generalized by Schanuel's conjecture.


Naming convention

The theorem is also known variously as the Hermite–Lindemann theorem and the Hermite–Lindemann–Weierstrass theorem.
Charles Hermite Charles Hermite () FRS FRSE MIAS (24 December 1822 – 14 January 1901) was a French mathematician who did research concerning number theory, quadratic forms, invariant theory, orthogonal polynomials, elliptic functions, and algebra. ...
first proved the simpler theorem where the exponents are required to be
rational integer An integer is the number zero (), a positive natural number (, , , etc.) or a negative integer with a minus sign ( −1, −2, −3, etc.). The negative numbers are the additive inverses of the corresponding positive numbers. In the language ...
s and linear independence is only assured over the rational integers, a result sometimes referred to as Hermite's theorem. Although apparently a rather special case of the above theorem, the general result can be reduced to this simpler case. Lindemann was the first to allow algebraic numbers into Hermite's work in 1882., . Shortly afterwards Weierstrass obtained the full result,, and further simplifications have been made by several mathematicians, most notably by
David Hilbert David Hilbert (; ; 23 January 1862 – 14 February 1943) was a German mathematician, one of the most influential mathematicians of the 19th and early 20th centuries. Hilbert discovered and developed a broad range of fundamental ideas in many ...
and
Paul Gordan __NOTOC__ Paul Albert Gordan (27 April 1837 – 21 December 1912) was a Jewish-German mathematician, a student of Carl Jacobi at the University of Königsberg before obtaining his PhD at the University of Breslau (1862),. and a professor ...
.


Transcendence of and

The
transcendence Transcendence, transcendent, or transcendental may refer to: Mathematics * Transcendental number, a number that is not the root of any polynomial with rational coefficients * Algebraic element or transcendental element, an element of a field exten ...
of and are direct corollaries of this theorem. Suppose is a non-zero algebraic number; then is a linearly independent set over the rationals, and therefore by the first formulation of the theorem is an algebraically independent set; or in other words is transcendental. In particular, is transcendental. (A more elementary proof that is transcendental is outlined in the article on
transcendental number In mathematics, a transcendental number is a number that is not algebraic—that is, not the root of a non-zero polynomial of finite degree with rational coefficients. The best known transcendental numbers are and . Though only a few classes ...
s.) Alternatively, by the second formulation of the theorem, if is a non-zero algebraic number, then is a set of distinct algebraic numbers, and so the set is linearly independent over the algebraic numbers and in particular cannot be algebraic and so it is transcendental. To prove that is transcendental, we prove that it is not algebraic. If were algebraic, ''i'' would be algebraic as well, and then by the Lindemann–Weierstrass theorem (see
Euler's identity In mathematics, Euler's identity (also known as Euler's equation) is the equality e^ + 1 = 0 where : is Euler's number, the base of natural logarithms, : is the imaginary unit, which by definition satisfies , and : is pi, the ratio of the circ ...
) would be transcendental, a contradiction. Therefore is not algebraic, which means that it is transcendental. A slight variant on the same proof will show that if is a non-zero algebraic number then and their hyperbolic counterparts are also transcendental.


-adic conjecture


Modular conjecture

An analogue of the theorem involving the modular function was conjectured by Daniel Bertrand in 1997, and remains an open problem. Writing for the square of the nome and the conjecture is as follows.


Lindemann–Weierstrass theorem


Proof

The proof relies on two preliminary lemmas. Notice that Lemma B itself is already sufficient to deduce the original statement of Lindemann–Weierstrass theorem.


Preliminary lemmas

Proof of Lemma A. To simplify the notation set: : \begin & n_0 =0, & & \\ & n_i =\sum\nolimits_^i m(k), & & i=1,\ldots,r \\ & n=n_r, & & \\ & \alpha_ =\gamma(i)_j, & & 1\leq i\leq r,\ 1\leq j\leq m(i) \\ & \beta_ =c(i). \end Then the statement becomes :\sum_^n \beta_k e^\neq 0. Let be a
prime number A prime number (or a prime) is a natural number greater than 1 that is not a Product (mathematics), product of two smaller natural numbers. A natural number greater than 1 that is not prime is called a composite number. For example, 5 is prime ...
and define the following polynomials: : f_i(x) = \frac , where is a non-zero integer such that \ell\alpha_1,\ldots,\ell\alpha_n are all algebraic integers. DefineUp to a factor, this is the same integral appearing in the proof that is a transcendental number, where The rest of the proof of the Lemma is analog to that proof. : I_i(s) = \int^s_0 e^ f_i(x) \, dx. Using
integration by parts In calculus, and more generally in mathematical analysis, integration by parts or partial integration is a process that finds the integral of a product of functions in terms of the integral of the product of their derivative and antiderivat ...
we arrive at : I_i(s) = e^s \sum_^ f_i^(0) - \sum_^ f_i^(s), where np-1 is the degree of f_i, and f_i^ is the ''j''-th derivative of f_i. This also holds for ''s'' complex (in this case the integral has to be intended as a contour integral, for example along the straight segment from 0 to ''s'') because :-e^ \sum_^ f_i^(x) is a primitive of e^ f_i(x). Consider the following sum: :\begin J_i &=\sum_^n\beta_k I_i(\alpha_k)\\ pt&= \sum_^n\beta_k \left ( e^ \sum_^ f_i^(0) - \sum_^ f_i^(\alpha_k)\right ) \\ pt&=\left(\sum_^f_i^(0)\right)\left(\sum_^n \beta_k e^\right)-\sum_^n\sum_^ \beta_kf_i^(\alpha_k)\\ pt&= -\sum_^n \sum_^ \beta_kf_i^(\alpha_k) \end In the last line we assumed that the conclusion of the Lemma is false. In order to complete the proof we need to reach a contradiction. We will do so by estimating , J_1\cdots J_n, in two different ways. First f_i^(\alpha_k) is an algebraic integer which is divisible by ''p''! for j\geq p and vanishes for j unless j=p-1 and k=i, in which case it equals :\ell^(p-1)!\prod_(\alpha_i-\alpha_k)^p. This is not divisible by ''p'' when ''p'' is large enough because otherwise, putting :\delta_i=\prod_(\ell\alpha_i-\ell\alpha_k) (which is a non-zero algebraic integer) and calling d_i\in\mathbb Z the product of its conjugates (which is still non-zero), we would get that ''p'' divides \ell^p(p-1)!d_i^p, which is false. So J_i is a non-zero algebraic integer divisible by (''p'' − 1)!. Now :J_i=-\sum_^\sum_^r c(t)\left(f_i^(\alpha_) + \cdots + f_i^(\alpha_)\right). Since each f_i(x) is obtained by dividing a fixed polynomial with integer coefficients by (x-\alpha_i), it is of the form :f_i(x)=\sum_^g_m(\alpha_i)x^m, where g_m is a polynomial (with integer coefficients) independent of ''i''. The same holds for the derivatives f_i^(x). Hence, by the fundamental theorem of symmetric polynomials, :f_i^(\alpha_)+\cdots+f_i^(\alpha_) is a fixed polynomial with rational coefficients evaluated in \alpha_i (this is seen by grouping the same powers of \alpha_,\dots,\alpha_ appearing in the expansion and using the fact that these algebraic numbers are a complete set of conjugates). So the same is true of J_i, i.e. it equals G(\alpha_i), where ''G'' is a polynomial with rational coefficients independent of ''i''. Finally J_1\cdots J_n=G(\alpha_1)\cdots G(\alpha_n) is rational (again by the fundamental theorem of symmetric polynomials) and is a non-zero algebraic integer divisible by (p-1)!^n (since the J_i's are algebraic integers divisible by (p-1)!). Therefore :, J_1\cdots J_n, \geq (p-1)!^n. However one clearly has: :, I_i(\alpha_k), \leq , \alpha_k, e^F_i(, \alpha_k, ), where is the polynomial whose coefficients are the absolute values of those of ''f''''i'' (this follows directly from the definition of I_i(s)). Thus :, J_i, \leq \sum_^n \left , \beta_k\alpha_k \right , e^F_i \left ( \left , \alpha_k \ \right ) and so by the construction of the f_i's we have , J_1\cdots J_n, \le C^p for a sufficiently large ''C'' independent of ''p'', which contradicts the previous inequality. This proves Lemma A. ∎ Proof of Lemma B: Assuming :b(1)e^+\cdots+ b(n)e^= 0, we will derive a contradiction, thus proving Lemma B. Let us choose a polynomial with integer coefficients which vanishes on all the \gamma(k)'s and let \gamma(1),\ldots,\gamma(n),\gamma(n+1),\ldots,\gamma(N) be all its distinct roots. Let ''b''(''n'' + 1) = ... = ''b''(''N'') = 0. The polynomial :P(x_1,\dots,x_N)=\prod_(b(1) x_+\cdots+b(N) x_) vanishes at (e^,\dots,e^) by assumption. Since the product is symmetric, for any \tau\in S_N the monomials x_^\cdots x_^ and x_1^\cdots x_N^ have the same coefficient in the expansion of ''P''. Thus, expanding P(e^,\dots,e^) accordingly and grouping the terms with the same exponent, we see that the resulting exponents h_1\gamma(1)+\dots+h_N\gamma(N) form a complete set of conjugates and, if two terms have conjugate exponents, they are multiplied by the same coefficient. So we are in the situation of Lemma A. To reach a contradiction it suffices to see that at least one of the coefficients is non-zero. This is seen by equipping with the lexicographic order and by choosing for each factor in the product the term with non-zero coefficient which has maximum exponent according to this ordering: the product of these terms has non-zero coefficient in the expansion and does not get simplified by any other term. This proves Lemma B. ∎


Final step

We turn now to prove the theorem: Let ''a''(1), ..., ''a''(''n'') be non-zero
algebraic number An algebraic number is a number that is a root of a non-zero polynomial in one variable with integer (or, equivalently, rational) coefficients. For example, the golden ratio, (1 + \sqrt)/2, is an algebraic number, because it is a root of th ...
s, and ''α''(1), ..., ''α''(''n'') distinct algebraic numbers. Then let us assume that: : a(1)e^+\cdots + a(n)e^ = 0. We will show that this leads to contradiction and thus prove the theorem. The proof is very similar to that of Lemma B, except that this time the choices are made over the ''a''(''i'')'s: For every ''i'' ∈ , ''a''(''i'') is algebraic, so it is a root of an irreducible polynomial with integer coefficients of degree ''d''(''i''). Let us denote the distinct roots of this polynomial ''a''(''i'')1, ..., ''a''(''i'')''d''(''i''), with ''a''(''i'')1 = ''a''(''i''). Let S be the functions σ which choose one element from each of the sequences (1, ..., ''d''(1)), (1, ..., ''d''(2)), ..., (1, ..., ''d''(''n'')), so that for every 1 ≤ ''i'' ≤ ''n'', σ(''i'') is an integer between 1 and ''d''(''i''). We form the polynomial in the variables x_,\dots,x_,\dots,x_,\dots,x_,y_1,\dots,y_n : Q(x_,\dots,x_,y_1,\dots,y_n)=\prod\nolimits_\left(x_y_1+\dots+x_y_n\right). Since the product is over all the possible choice functions σ, ''Q'' is symmetric in x_,\dots,x_ for every ''i''. Therefore ''Q'' is a polynomial with integer coefficients in elementary symmetric polynomials of the above variables, for every ''i'', and in the variables ''y''''i''. Each of the latter symmetric polynomials is a rational number when evaluated in a(i)_1,\dots,a(i)_. The evaluated polynomial Q(a(1)_1,\dots,a(n)_,e^,\dots,e^) vanishes because one of the choices is just σ(''i'') = 1 for all ''i'', for which the corresponding factor vanishes according to our assumption above. Thus, the evaluated polynomial is a sum of the form : b(1)e^+ b(2)e^+ \cdots + b(N)e^= 0, where we already grouped the terms with the same exponent. So in the left-hand side we have distinct values β(1), ..., β(''N''), each of which is still algebraic (being a sum of algebraic numbers) and coefficients b(1),\dots,b(N)\in\mathbb Q. The sum is nontrivial: if \alpha(i) is maximal in the lexicographic order, the coefficient of e^ is just a product of ''a''(''i'')''j'''s (with possible repetitions), which is non-zero. By multiplying the equation with an appropriate integer factor, we get an identical equation except that now ''b''(1), ..., ''b''(''N'') are all integers. Therefore, according to Lemma B, the equality cannot hold, and we are led to a contradiction which completes the proof. ∎ Note that Lemma A is sufficient to prove that ''e'' is
irrational Irrationality is cognition, thinking, talking, or acting without inclusion of rationality. It is more specifically described as an action or opinion given through inadequate use of reason, or through emotional distress or cognitive deficiency. T ...
, since otherwise we may write ''e'' = ''p'' / ''q'', where both ''p'' and ''q'' are non-zero integers, but by Lemma A we would have ''qe'' − ''p'' ≠ 0, which is a contradiction. Lemma A also suffices to prove that is irrational, since otherwise we may write = ''k'' / ''n'', where both ''k'' and ''n'' are integers) and then ±''i'' are the roots of ''n''2''x''2 + ''k''2 = 0; thus 2 − 1 − 1 = 2''e''0 + ''e''''i'' + ''e''−''i'' ≠ 0; but this is false. Similarly, Lemma B is sufficient to prove that ''e'' is transcendental, since Lemma B says that if ''a''0, ..., ''a''''n'' are integers not all of which are zero, then : a_ne^n+\cdots+a_0e^0\ne 0. Lemma B also suffices to prove that is transcendental, since otherwise we would have 1 + ''e''''i'' ≠ 0.


Equivalence of the two statements

Baker's formulation of the theorem clearly implies the first formulation. Indeed, if \alpha(1),\ldots,\alpha(n) are algebraic numbers that are linearly independent over \Q, and :P(x_1, \ldots, x_n)= \sum b_ x_1^\cdots x_n^ is a polynomial with rational coefficients, then we have :P\left(e^,\dots,e^\right)= \sum b_ e^, and since \alpha(1),\ldots,\alpha(n) are algebraic numbers which are linearly independent over the rationals, the numbers i_1 \alpha(1) + \cdots + i_n \alpha(n) are algebraic and they are distinct for distinct ''n''-tuples (i_1,\dots,i_n). So from Baker's formulation of the theorem we get b_=0 for all ''n''-tuples (i_1,\dots,i_n). Now assume that the first formulation of the theorem holds. For n=1 Baker's formulation is trivial, so let us assume that n>1, and let a(1),\ldots,a(n) be non-zero algebraic numbers, and \alpha(1),\ldots,\alpha(n) distinct algebraic numbers such that: :a(1)e^ + \cdots + a(n)e^ = 0. As seen in the previous section, and with the same notation used there, the value of the polynomial :Q(x_,\ldots,x_,y_1,\dots,y_n)=\prod\nolimits_\left(x_y_1+\dots+x_y_n\right), at :\left (a(1)_1,\ldots,a(n)_,e^,\ldots,e^ \right) has an expression of the form : b(1)e^+ b(2)e^+ \cdots + b(M)e^= 0, where we have grouped the exponentials having the same exponent. Here, as proved above, b(1),\ldots, b(M) are rational numbers, not all equal to zero, and each exponent \beta(m) is a linear combination of \alpha(i) with integer coefficients. Then, since n>1 and \alpha(1),\ldots,\alpha(n) are pairwise distinct, the \Q-vector subspace V of \C generated by \alpha(1),\ldots,\alpha(n) is not trivial and we can pick \alpha(i_1),\ldots,\alpha(i_k) to form a basis for V. For each m=1,\dots,M, we have :\begin \beta(m) = q_ \alpha(i_1) + \cdots + q_ \alpha(i_k), && q_ = \frac; \qquad c_, d_ \in \Z. \end For each j=1,\ldots, k, let d_j be the least common multiple of all the d_ for m=1,\ldots, M, and put v_j = \tfrac \alpha(i_j). Then v_1,\ldots,v_k are algebraic numbers, they form a basis of V, and each \beta(m) is a linear combination of the v_j with integer coefficients. By multiplying the relation :b(1)e^+ b(2)e^+ \cdots + b(M)e^= 0, by e^, where N is a large enough positive integer, we get a non-trivial algebraic relation with rational coefficients connecting e^,\cdots,e^, against the first formulation of the theorem.


See also

*
Gelfond–Schneider theorem In mathematics, the Gelfond–Schneider theorem establishes the transcendence of a large class of numbers. History It was originally proved independently in 1934 by Aleksandr Gelfond and Theodor Schneider. Statement : If ''a'' and ''b'' a ...
*
Baker's theorem In transcendental number theory, a mathematical discipline, Baker's theorem gives a lower bound for the absolute value of linear combinations of logarithms of algebraic numbers. The result, proved by , subsumed many earlier results in transcendent ...
; an extension of Gelfond–Schneider theorem * Schanuel's conjecture; if proven, it would imply both the Gelfond–Schneider theorem and the Lindemann–Weierstrass theorem


Notes


References

* * * * * * * *


Further reading

* * * *


External links

* * {{DEFAULTSORT:Lindemann-Weierstrass theorem Articles containing proofs E (mathematical constant) Exponentials Pi Theorems in number theory Transcendental numbers