Fact-checked by Grok 2 weeks ago

Auxiliary function

In , particularly in , an auxiliary function is a specially constructed —often a , , or more complex analytic expression—designed to facilitate proofs of , , or of numbers such as e, \pi, or values of the . These functions are engineered to exhibit controlled behavior, such as having small values or specific zeros at algebraic points of interest, which enables the application of inequalities like or the to derive contradictions or precise bounds that confirm the transcendental nature of the target quantities. The development of auxiliary functions traces back to the mid-19th century, beginning with Joseph Liouville's 1844 construction of polynomials to demonstrate the transcendence of certain infinite series, such as \sum 10^{-n!}, marking the first explicit proof of a transcendental number. Charles Hermite advanced this in 1873 by introducing Padé approximants as auxiliary functions to prove the transcendence of e, using forms like B(z)e^z - A(z) where A(z) and B(z) are polynomials tailored to minimize discrepancies at integer points. Subsequent refinements by mathematicians including Axel Thue, Carl Ludwig Siegel, Aleksandr Gelfond, and Theodor Schneider in the early 20th century incorporated tools like the Thue–Siegel lemma (based on Dirichlet's box principle) to handle multivariable cases, enabling proofs for numbers like e^\pi and \pi. By the mid-20th century, Kurt Mahler introduced linear algebra techniques for constructing to solve functional equations, while later innovations, such as Michel Laurent's 1991 use of interpolation determinants, eliminated reliance on for more efficient constructions. These functions have been in addressing major open problems, including Hilbert's seventh problem on the of a^b for algebraic a \neq 0,1 and irrational algebraic b, and continue to underpin modern results in , such as for values of the at rational arguments.

Fundamentals

Definition

In , which concerns numbers that are not roots of any non-zero equation with rational coefficients, auxiliary functions serve as essential tools for proving or . These functions address the fundamental challenge of distinguishing algebraic numbers—those satisfying such equations—from by constructing approximations that algebraic numbers cannot achieve due to limitations imposed by their and . Specifically, transcendence proofs often rely on showing that certain numbers can be approximated by rationals or algebraics to an extent that violates known bounds for algebraic numbers, such as or Liouville's inequality extensions. An auxiliary function is a specially constructed analytic or , typically a , rational approximant, or integral expression, engineered to exhibit particular boundedness properties, such as taking exceptionally small values at rational or algebraic points of interest while remaining non-zero overall. The core purpose is to derive a : if the number in question were algebraic, the function's smallness at those points would imply it vanishes identically or equals zero, which it does not, thus establishing . These functions are derived using techniques like the , linear algebra (e.g., Thue-Siegel lemma), or to ensure the desired estimates. Common general forms include sums like \sum_{k=0}^m a_k x^k / k! where x is algebraic and coefficients a_k are chosen rationally. properties of auxiliary functions encompass their non-integral at algebraic points (ensuring they are not integers or algebraic integers in the assumed case), strict positivity in many analytic constructions to avoid sign changes, and established lower bounds on their minima or norms, which prevent trivial zero behavior and support the . These properties are rigorously controlled through estimates on , growth rates, or zero multiplicities at specified points.

Historical Development

The origins of auxiliary functions in mathematical proofs trace back to Joseph Liouville's pioneering work, with his approximation theorem in 1844 laying groundwork through to show that certain real numbers are not algebraic. In his seminal 1851 memoir "Sur des classes très-étendues de quantités dont la valeur n'est ni algébrique, ni même réductible à des irrationnelles algébriques," published in the Journal de Mathématiques Pures et Appliquées, Liouville employed these functions to construct numbers with exceptionally good rational approximations, thereby proving their and laying the groundwork for the field. Advancements in the built on this foundation, with Joseph Fourier's 1815 argument for the of e—published posthumously—utilizing series to generate sharp rational approximations, serving as an early precursor to auxiliary function techniques in and proofs. Hermite advanced the method significantly in 1873 by proving the of e through the construction of an explicit auxiliary function involving integrals of exponential terms, which demonstrated that e satisfies no with rational coefficients; this memoir in the Comptes Rendus de l'Académie des Sciences marked the first such proof for a specific . In the 20th century, extended Hermite's approach in 1882 to prove the of π by showing that e^{iπ} = -1 implies π's non-algebraicity via auxiliary functions approximating exponential integrals. This culminated in the 1934 Gelfond-Schneider theorem, independently proved by Aleksandr Gelfond and Theodor Schneider, which established the transcendence of numbers of the form α^β for algebraic α ≠ 0,1 and irrational algebraic β, using auxiliary polynomials in entire functions like e^z and e^{βz}. The evolution of auxiliary functions shifted from these explicit, integral-based constructions in the to more abstract formulations by the mid-20th century, incorporating the for bounds and methods for approximating transcendental functions with algebraic ones.

Explicit Constructions

Liouville's Transcendence Criterion

Liouville's transcendence criterion, published in 1844, establishes a sufficient condition for a \alpha to be transcendental based on the quality of its rational approximations. If \alpha is algebraic of degree n, then there exists a positive constant c (depending on \alpha) such that |\alpha - p/q| > c / q^n for all integers p, q with q > 0. Thus, if there are infinitely many rationals p/q satisfying |\alpha - p/q| < 1/q^\kappa for some \kappa > n, then \alpha cannot be algebraic and must be transcendental. The proof relies on constructing an auxiliary function from the minimal polynomial of \alpha. Assume \alpha is algebraic of degree n with irreducible minimal polynomial f(x) \in \mathbb{Z} of degree n, so f(\alpha) = 0. For a rational p/q in lowest terms, q^n f(p/q) is a non-zero integer because f is irreducible over \mathbb{Q}, implying |f(p/q)| \geq 1/q^n. This polynomial f serves as the auxiliary function. To derive the approximation bound, apply the : |f(\alpha) - f(p/q)| \leq \sup |f'(x)| \cdot |\alpha - p/q| over the interval between \alpha and p/q, where the supremum is bounded by some constant M depending on f and the interval. Since f(\alpha) = 0, it follows that |f(p/q)| \leq M |\alpha - p/q|, so |\alpha - p/q| \geq 1/(M q^n). If approximations better than this bound exist infinitely often (i.e., with exponent \kappa > n), the assumption that \alpha is algebraic leads to a . The key is thus |f(p/q)| > c / q^\mu for \mu = n and some c > 0, which cannot hold if |\alpha - p/q| is sufficiently small. A seminal application is to Liouville's constant \alpha = \sum_{k=1}^\infty 10^{-k!}, the first explicit constructed in 1844. Consider the partial sum up to m, p/q = \sum_{k=1}^m 10^{-k!} with q = 10^{m!}, so p, q are integers. The tail satisfies |\alpha - p/q| = \sum_{k=m+1}^\infty 10^{-k!} < 10^{-(m+1)!} / (1 - 10^{-(m+1)!}) < 2 \cdot 10^{-(m+1)!} for m \geq 2. Since (m+1)! = (m+1) \cdot m!, this is |\alpha - p/q| < 2 / q^{m+1}. For fixed degree n, choose m > n; then \kappa = m+1 > n, yielding infinitely many such approximations as m increases, so \alpha is transcendental by the criterion using the auxiliary minimal polynomial.

Fourier's Proof of the Irrationality of e

outlined a proof of the of e in a dated around 1815, which was communicated by Poinsot and published the same year in Janot de Stainville's Mélanges d'analyse algébrique et de géométrie. This proof represents the first rigorous demonstration that e cannot be expressed as a of integers, predating more advanced results on its . The proof centers on the Taylor series expansion of e: e = \sum_{k=0}^{\infty} \frac{1}{k!}. The auxiliary function is the remainder term after truncating the series at n terms, R_n = e - \sum_{k=0}^{n} \frac{1}{k!} = \sum_{k=n+1}^{\infty} \frac{1}{k!}, which captures the tail of the series and satisfies $0 < R_n < \frac{1}{n \cdot n!}. This remainder can equivalently be expressed in integral form to aid in bounding its value, R_n = \int_{0}^{1} \frac{(1-t)^n}{n!} e^{t} \, dt, derived from the integral form of the Taylor remainder theorem applied to e^x at x=1. The integral representation highlights the positive nature and upper bound of the tail, with R_n > \int_{0}^{1} \frac{(1-t)^n}{n!} \, dt = \frac{1}{(n+1)!} and R_n < \frac{e}{(n+1)!}, ensuring $0 < n! R_n < \frac{e}{n+1} < 1 for sufficiently large n. To establish irrationality, assume for contradiction that e = a/b where a and b are positive integers with \gcd(a,b)=1. Set n = b. Then b! e = a \cdot (b-1)!, which is an integer. Alternatively, expanding the series gives b! e = \sum_{k=0}^{b} \frac{b!}{k!} + b! R_b, where \sum_{k=0}^{b} \frac{b!}{k!} is an integer because k! divides b! for k \leq b. Thus, b! R_b must also be an integer. However, the bounds on the remainder imply \frac{1}{b+1} < b! R_b < \frac{e}{b+1} < 1 for b \geq 3, so $0 < b! R_b < 1, contradicting the assumption that it is a nonzero integer. Therefore, e is irrational. A key aspect of the proof involves scaling the remainder via integrals related to the gamma function, where the full integral satisfies \int_{0}^{\infty} e^{-t} \frac{t^n}{n!} \, dt = 1, providing context for the tail's magnitude when truncated and scaled appropriately to yield the strict inequality $0 < b! R_b < 1. This approach using the series tail as an auxiliary function marks the earliest known rigorous argument for e's irrationality, relying on the exponential series without invoking continued fractions or other methods.

Hermite's Proofs Involving e

In 1873, Charles Hermite extended earlier work on the irrationality of e, such as Fourier's 1815 proof using integral remainders of the exponential series, to establish the irrationality of e^r for any nonzero rational r and the transcendence of e itself. His contributions appeared in a series of notes in the Comptes rendus hebdomadaires des séances de l'Académie des sciences and were consolidated in a memoir published in Journal de mathématiques pures et appliquées (Liouville's Journal). These proofs relied on explicit constructions of auxiliary functions via Padé approximants and integrals, which allowed Hermite to derive contradictions from assumptions of algebraicity by bounding the functions' values at integers while preserving integrality properties. For the irrationality of e^r where r is a nonzero rational, Hermite constructed auxiliary functions as rational approximations to e^z, specifically polynomials A(z) and B(z) with integer coefficients such that the auxiliary R(z) = B(z) e^z - A(z) has a zero of high multiplicity at z=0. Evaluating at z=r yields B(r) e^r - A(r) = R(r), where $0 < |R(r)| < 1 for large multiplicity after appropriate scaling (e.g., multiplying by a denominator to make A(r), B(r) integers), contradicting the assumption that e^r = a/b rational, as it would imply R(r) = 0. This approach generalized Fourier's integral-based remainder estimates for the exponential function. To prove the transcendence of e, Hermite extended these constructions using auxiliary polynomials f(t) with high-order zeros at integers t=0,1,\dots,m, such as f_r(t) = t^{r-1} (t-1)^r \cdots (t-m)^r, and integrals I_k = \int_0^k e^{-t} f(t) \, dt for k=0,1,\dots,m. Under the assumption that e satisfies a polynomial equation \sum_{j=0}^d a_j e^j = 0 with a_j \in \mathbb{Z}, a_d \neq 0, a linear combination \Phi = \sum_{j=0}^d a_j e^j I_j (adjusted via Hermite's identity relating integrals and polynomials) is a nonzero . However, bounds show $0 < |\Phi| < 1 for sufficiently large r > d, using factorial decay in the integrals, leading to a contradiction since no such small nonzero exists. This integral satisfies properties ensuring no exact cancellation, ruling out algebraic relations. These methods marked a pivotal advance, introducing systematic use of auxiliary integrals and Padé approximants for Diophantine approximations in .

Pigeonhole Principle Applications

Auxiliary Polynomial Theorem

In , auxiliary polynomials are constructed using the to obtain non-zero polynomials with coefficients that take exceptionally small values at a given \alpha of degree d over \mathbb{Q} and bounded . These polynomials enable effective lower bounds on how well \alpha can be approximated by . The construction often begins by applying the in the unit to vectors of evaluations (P(0), P(1), \dots, P(m)) modulo 1 in [0,1]^{m+1}, where m is on the order of d. By considering a large collection of polynomials with bounded coefficients, their evaluation vectors fill the . With more vectors than subintervals of a suitable grid, the principle guarantees two distinct polynomials whose difference P (non-zero) has evaluation vector components differing by less than the grid spacing, yielding small fractional parts and thus small |P(k)| for k = 0 to m. This framework, foundational to the Thue-Siegel method, extends to the conjugates of \alpha using tools like Siegel's lemma—a pigeonhole-based result on small solutions to linear systems—to achieve simultaneous smallness |P(\beta_i)| < \varepsilon for each Galois conjugate \beta_i of \alpha (i = 1, \dots, d), where \varepsilon is controlled by the grid size $1/N and N exceeds the number of candidates. This uniform control ensures the polynomial is non-trivial while providing quantitative estimates for contradiction arguments in approximation theorems. The degree n and height of such P are balanced to make |P(\alpha)| small relative to the height H(P), typically achieving |P(\alpha)| \ll H(P)^{-\mu} for some \mu > 1 depending on d, which underpins measures for algebraic numbers. The height Q of \alpha, reflecting the coefficients of its minimal , influences the scale of the . This approach serves as a precursor to advanced results like on the irrationality measure of algebraic numbers, by enabling controlled constructions of polynomials small near algebraic points.

Lang's Theorem on Diophantine Approximation

The Schneider-Lang theorem on , a refinement by of Theodor Schneider's work in the , provides a powerful criterion for limiting the algebraic values taken by s of finite , with significant implications for . Specifically, consider meromorphic functions f_1, \dots, f_m in \mathbb{C} of finite order, where f_1 and f_2 are algebraically over \mathbb{Q}(z), and the derivatives satisfy f_j'(w) \in K(f_1(w), \dots, f_m(w)) for a number field K. Then, the set S = \{ w \in \mathbb{C} \mid w is not a pole of any f_j, and f_j(w) \in K for all j = 1, \dots, m \} is finite. In the case of a single f(z) of finite order \rho, this implies that there are only finitely many rationals p/q (in lowest terms) such that |f(p/q)| < 1/|q|^\kappa for any \kappa > \rho, unless f is a special satisfying an algebraic over \mathbb{Q}(z). The proof relies on constructing auxiliary functions via the to exploit the growth properties of entire functions. One key step involves Dirichlet's applied to points in the space of integer linear combinations of basis functions derived from the f_j. This allows selection of non-trivial integer coefficients b_1, \dots, b_n such that the auxiliary function g(z) = \sum b_i \prod f_j(z)^{a_{ij}} (or a similar form) vanishes at many points in S \cap D(0, R), where D(0, R) is a disk of radius R. For applications involving approximations, auxiliary functions of the form \exp(g(z)), where g(z) is a with algebraic coefficients, are used to bound distances like |f(\alpha) - \beta| for algebraic \alpha, \beta \in K, leveraging the finite order to control growth outside the disk. In the proof, after constructing g(z) to have at least N zeros within a smaller disk |z| \leq r < R, analytic estimates such as the provide an upper bound: |g(0)| \leq \left( \frac{3r}{R} \right)^N \max_{|z|=R} |g(z)|. A lower bound for |g(0)| is then obtained via on the coefficients, ensuring that if S were infinite, the growth would contradict the finite order unless the functions are dependent. This builds on the by extending algebraic pigeonhole arguments to analytic settings. A primary application lies in Baker's method for lower bounds on linear forms in logarithms, where Lang's framework refines estimates for forms \Lambda = b_0 + b_1 \log \alpha_1 + \dots + b_n \log \alpha_n with algebraic \alpha_i and integer b_i. Using auxiliary polynomials in the exponents, the method yields |\Lambda| > \exp(-C (\log H)^\tau), where H is the (exponential) height of the \alpha_i and b_i, and C, \tau > 0 depend on the degree and n. This quantitative result, pivotal for solving Diophantine equations, stems directly from the Diophantine control imposed by the theorem on approximations near algebraic points.

Interpolation Methods

General Interpolation Determinants

In transcendence theory, auxiliary functions are often constructed using interpolation determinants, which are scalars derived from matrices whose entries involve evaluations of a function and its derivatives at specified points. Specifically, for an f and distinct points a_1, \dots, a_n, the interpolation matrix M has entries M_{ij} = f^{(i-1)}(a_j) / (i-1)! for i, j = 1, \dots, n, and the auxiliary is the \det(M). This construction generalizes the Vandermonde determinant from to analytic functions, providing a measure of (or near-dependence) among the vectors of function values and scaled derivatives at the points a_j. A small |\det(M)| indicates that f is well-approximated by a of degree less than n at these points, which is key for deriving upper bounds in transcendence arguments. Upper bounds on the magnitude of \det(M) are obtained using , which limits the growth based on the row norms, often combined with analytic estimates like the to bound derivatives within a disk of . For instance, if f is entire and bounded on a disk of R, the determinant satisfies |\det(M)| \leq n^{n/2} \prod_{j=1}^n \max_{0 \leq k < n} |f^{(k)}(a_j)/k!|. Lower bounds in generic cases ensure non-vanishing, with estimates such as |\det(M)| > \exp(-n^2 \log n) for suitably chosen points. These properties enable tight control over approximation quality, crucial for contradictions in transcendence proofs. The method traces its historical roots to the late , stemming from Weierstrass's work on entire functions and techniques to bound zeros. It was refined in the , notably by Schneider, and formalized by M. Laurent in 1991 through explicit constructions for exponential polynomials that avoid the . A foundational example is the Vandermonde \Delta = \det((x_j^{i-1})_{1 \leq i,j \leq n}), which for points x_j related to exponential bases (e.g., x_j \approx e^{\alpha_j}) and |x_j| \leq 1 admits lower bounds such as |\Delta| > \exp(-n^2 \log n) from product formulas and , ensuring non-vanishing in generic configurations. These interpolation determinants bridge classical interpolation with analytic estimates, facilitating global transcendence results from local approximations. A prominent application is in proofs of the Hermite–Lindemann theorem, where they construct linear forms in exponential values at algebraic points to establish transcendence.

Proof of the Hermite–Lindemann Theorem

The Hermite–Lindemann theorem asserts that if \alpha is a non-zero algebraic number, then e^{\alpha} is transcendental. A key consequence is the transcendence of \pi, since e^{i\pi} = -1 and i\pi is algebraic, so assuming \pi algebraic would imply e^{i\pi} algebraic, contradicting the theorem. Lindemann's 1882 proof extends Hermite's 1873 demonstration of the transcendence of e by generalizing the auxiliary function approach to algebraic exponents. To establish transcendence, assume for contradiction that e^{\alpha} is algebraic, where \alpha \neq 0 is algebraic of degree d. Consider the field extension \mathbb{Q}(\alpha, e^{\alpha}) of degree at most d^2, and let \alpha_1 = \alpha, \dots, \alpha_m be the conjugates of \alpha under the , with m \leq d. The assumption implies e^{\alpha_j} are algebraic for all j. The core of the proof relies on constructing an auxiliary function via determinants to capture linear relations among the e^{\alpha_j}. Specifically, the coefficients b_1, \dots, b_m are defined using determinants of matrices that enforce conditions on the exponentials at points. This yields the auxiliary function \phi(z) = \sum_{j=1}^m b_j e^{\alpha_j z}, which satisfies \phi(k) = 0 for k = 1, \dots, n (with n large) under the linear dependence assumption derived from the algebraic hypothesis. Under the , \phi(z) is non-zero (as the e^{\alpha_j z} are linearly over \overline{\mathbb{Q}} in generic settings), and \phi(0) = \sum b_j is a non-zero element of the number field with bounded denominator (from the field properties and coefficients in the determinants), implying |\phi(0)| \geq c > 0 for some positive constant c of n. However, estimates on \phi(z) using the multiple zeros at k=1,\dots,n, combined with on the coefficient determinants and analytic bounds on the exponentials, yield $0 < |\phi(0)| < \exp(-c n^2) for large n and suitable c > 0. This smallness arises from the conditions suppressing the values near the origin. The contradiction between the lower bound from algebraicity and the exponential upper bound proves the false. The auxiliary \phi(z) satisfies a of order m (the number of distinct \alpha_j), as it is a of solutions to DEs y' = \alpha_j y. Applying the Phragmén–Lindelöf principle to this on suitable contours ensures \phi(z) does not vanish identically and reinforces the lower bound at z=0. This differential structure, combined with interpolation properties, distinguishes the proof from earlier methods and solidifies the result.

References

  1. [1]
    [0908.4024] Auxiliary functions in transcendence proofs - arXiv
    Aug 27, 2009 · Abstract: We discuss the role of auxiliary functions in the development of transcendental number theory. Subjects: Number Theory (math.NT).
  2. [2]
    [PDF] Auxiliary functions in transcendence proofs - HAL
    Jul 24, 2009 · We discuss the role of auxiliary functions in the development of transcendental number theory. Earlier auxiliary functions were completely ...
  3. [3]
    [PDF] AWS Lecture 4 Auxiliary functions in transcendence proofs
    Transcendence proofs most often involve an auxiliary function. Such functions can take several forms. Historically, the first ones were Padé approximations, ...
  4. [4]
    Auxiliary functions in transcendental number theory
    Oct 24, 2009 · We discuss the role of auxiliary functions in the development of transcendental number theory. Initially, auxiliary functions were completely explicit.
  5. [5]
    [PDF] A Historical Introduction to Transcendental Number Theory
    Among the numbers conjectured to be transcendental after Liouville's results, e was the first to be proven so, by Charles Hermite in 1873. We begin our ...<|separator|>
  6. [6]
    [PDF] Construction of Transcendental Numbers
    The construction of a Transcendental number in the second part of these notes using a series is due to Joseph Liouville. We also point out that the great ...<|control11|><|separator|>
  7. [7]
    [PDF] IRRATIONALITY OF π AND e 1. Introduction Numerical estimates for ...
    In Section 2, we prove π is irrational with definite integrals. Irrationality of e is proved by infinite series in Section 3. A general discussion about ...
  8. [8]
    [PDF] Fourier's Infinite Series Proof of the Irrationality of e
    Oct 3, 2022 · number systems. Kenneth M Monks, “Fourier's Infinite Series Proof of the Irrationality of e”. MAA Convergence (October 2022).
  9. [9]
    [PDF] Irrationality proofs of e and π
    Our first proof is by Steve Kifowit from 2009. It starts by taking the definite integral of e−x from. 0 to 1 and applying integration by parts a variable, but ...
  10. [10]
  11. [11]
    [PDF] Irrationality proofs `a la Hermite - arXiv
    Aug 16, 2010 · Theorem 1. If r2 ∈ Q \ {0} then r tan r is irrational. Proof. The irrationality of π2 will be a byproduct of this proof ...
  12. [12]
    [PDF] TRANSCENDENCE OF e - KEITH CONRAD
    Complex numbers that are not algebraic are called transcendental. A transcendental number is a “strong” type of irrational number: the irrational numbers are.
  13. [13]
  14. [14]
    [PDF] transcendental number theory - Mathematics
    1. Liouville's theorem. The theory of transcendental numbers was originated by Liouville in his famous memoir of 1844 in which he obtained, for the ...Missing: Joseph | Show results with:Joseph
  15. [15]
    [PDF] Chapter 4 Transcendence results
    4.1 The transcendence of e. We define as usual ez = P. ∞ n=0 zn/n! for z ∈ C. Further, Q = {α ∈ C : α algebraic over Q}. Theorem 4.1 (Hermite, 1873). e is ...
  16. [16]
    [PDF] Michel Waldschmidt Abstract Schneider – Lang Theorem ... - IMJ-PRG
    Jan 15, 2016 · The proof of the Schneider – Lang Theorem follows the following scheme : Step 1 Construct an auxiliary function f with many zeroes. Step 2 Find ...
  17. [17]
    Introduction to Diophantine Approximations
    Lang, Serge, 1927-. Introduction to diophantine approximations / Serge Lang. p. cm. Originally published: Reading, Mass. : Addison-Wesley Pub. Co.,. 1966.
  18. [18]
    [PDF] Diophantine Approximation on Linear Algebraic Groups
    The Criterion of Schneider-Lang. 115. 4.1 Algebraic Values of Entire Functions Satisfying Differential. Equations. 115. 4.2 FirstProof of Baker's Theorem. 118.
  19. [19]
    On Baker's Method | SpringerLink
    In Chap. 4 we deduced Baker's Theorems 1.5 and 1.6 from Schneider-Lang's Criterion. The proof used an extension of Gel'fond's method in several variables.
  20. [20]
    [PDF] Diophantine Analysis, Michaelmas 2024 - University of Cambridge
    A determinant of this form is called an interpolation determinant. When there are no derivatives involved, it is called an alternant. So the proof can described ...
  21. [21]
    [PDF] Extrapolation with interpolation determinants
    Their proofs are different, but both involve the construction of an auxiliary function using lemma 1.1. Until recently, most proofs of transcendental number ...<|control11|><|separator|>
  22. [22]
    [PDF] AWS Lecture 2 Historical introduction to transcendence
    The transcendence proofs for constants of analysis are essentially all based on the seminal work by Ch. Hermite : his proof of the transcendence of the ...
  23. [23]
  24. [24]
    TRANSCENDENTAL NUMBER THEORY
    An auxiliary result in the theory of transcendental numbers, J. Number ... auxiliary function, 9, 10, 13-20, 24, 28-. 34, 58-60, 62, 64, 125, 126, 127.