Transcendental number

From Wikipedia, the free encyclopedia
(Redirected from Transcendental Number)

In mathematics, a transcendental number is a real or complex 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 π and e.[1][2]

Though only a few classes of transcendental numbers are known – partly because it can be extremely difficult to show that a given number is transcendental – transcendental numbers are not rare: indeed, almost all real and complex numbers are transcendental, since the algebraic numbers form a countable set, while the set of real numbers and the set of complex numbers are both uncountable sets, and therefore larger than any countable set. All transcendental real numbers (also known as real transcendental numbers or transcendental irrational numbers) are irrational numbers, since all rational numbers are algebraic.[3][4][5][6] The converse is not true: Not all irrational numbers are transcendental. Hence, the set of real numbers consists of non-overlapping sets of rational, algebraic non-rational, and transcendental real numbers.[3] For example, the square root of 2 is an irrational number, but it is not a transcendental number as it is a root of the polynomial equation x2 − 2 = 0. The golden ratio (denoted or ) is another irrational number that is not transcendental, as it is a root of the polynomial equation x2x − 1 = 0. The quality of a number being transcendental is called transcendence.

History[edit]

The name "transcendental" comes from Latin trānscendere 'to climb over or beyond, surmount',[7] and was first used for the mathematical concept in Leibniz's 1682 paper in which he proved that sin x is not an algebraic function of x .[8] Euler, in the 18th century, was probably the first person to define transcendental numbers in the modern sense.[9]

Johann Heinrich Lambert conjectured that e and π were both transcendental numbers in his 1768 paper proving the number π is irrational, and proposed a tentative sketch proof that π is transcendental.[10]

Joseph Liouville first proved the existence of transcendental numbers in 1844,[11] and in 1851 gave the first decimal examples such as the Liouville constant

in which the nth digit after the decimal point is 1 if n is equal to k! (k factorial) for some k and 0 otherwise.[12] In other words, the nth digit of this number is 1 only if n is one of the numbers 1! = 1, 2! = 2, 3! = 6, 4! = 24, etc. Liouville showed that this number belongs to a class of transcendental numbers that can be more closely approximated by rational numbers than can any irrational algebraic number, and this class of numbers are called Liouville numbers, named in his honour. Liouville showed that all Liouville numbers are transcendental.[13]

The first number to be proven transcendental without having been specifically constructed for the purpose of proving transcendental numbers' existence was e, by Charles Hermite in 1873.

In 1874, Georg Cantor proved that the algebraic numbers are countable and the real numbers are uncountable. He also gave a new method for constructing transcendental numbers.[14] Although this was already implied by his proof of the countability of the algebraic numbers, Cantor also published a construction that proves there are as many transcendental numbers as there are real numbers.[a] Cantor's work established the ubiquity of transcendental numbers.

In 1882, Ferdinand von Lindemann published the first complete proof that π is transcendental. He first proved that ea is transcendental if a is a non-zero algebraic number. Then, since e = −1 is algebraic (see Euler's identity), must be transcendental. But since i is algebraic, π must therefore be transcendental. This approach was generalized by Karl Weierstrass to what is now known as the Lindemann–Weierstrass theorem. That π is transcendental implies that geometric constructions involving compass and straightedge cannot produce certain results, for example squaring the circle.

In 1900, David Hilbert posed a question about transcendental numbers, Hilbert's seventh problem: If a is an algebraic number that is not zero or one, and b is an irrational algebraic number, is ab necessarily transcendental? The affirmative answer was provided in 1934 by the Gelfond–Schneider theorem. This work was extended by Alan Baker in the 1960s in his work on lower bounds for linear forms in any number of logarithms (of algebraic numbers).[16]

Properties[edit]

A transcendental number is a (possibly complex) number that is not the root of any integer polynomial. Every real transcendental number must also be irrational, since a rational number is the root of an integer polynomial of degree one.[17] The set of transcendental numbers is uncountably infinite. Since the polynomials with rational coefficients are countable, and since each such polynomial has a finite number of zeroes, the algebraic numbers must also be countable. However, Cantor's diagonal argument proves that the real numbers (and therefore also the complex numbers) are uncountable. Since the real numbers are the union of algebraic and transcendental numbers, it is impossible for both subsets to be countable. This makes the transcendental numbers uncountable.

No rational number is transcendental and all real transcendental numbers are irrational. The irrational numbers contain all the real transcendental numbers and a subset of the algebraic numbers, including the quadratic irrationals and other forms of algebraic irrationals.

Applying any non-constant single-variable algebraic function to a transcendental argument yields a transcendental value. For example, from knowing that π is transcendental, it can be immediately deduced that numbers such as , , , and are transcendental as well.

However, an algebraic function of several variables may yield an algebraic number when applied to transcendental numbers if these numbers are not algebraically independent. For example, π and (1 − π) are both transcendental, but π + (1 − π) = 1 is obviously not. It is unknown whether e + π, for example, is transcendental, though at least one of e + π and must be transcendental. More generally, for any two transcendental numbers a and b, at least one of a + b and ab must be transcendental. To see this, consider the polynomial (xa)(xb) = x2 − (a + b) x + a b . If (a + b) and a b were both algebraic, then this would be a polynomial with algebraic coefficients. Because algebraic numbers form an algebraically closed field, this would imply that the roots of the polynomial, a and b, must be algebraic. But this is a contradiction, and thus it must be the case that at least one of the coefficients is transcendental.

The non-computable numbers are a strict subset of the transcendental numbers.

All Liouville numbers are transcendental, but not vice versa. Any Liouville number must have unbounded partial quotients in its continued fraction expansion. Using a counting argument one can show that there exist transcendental numbers which have bounded partial quotients and hence are not Liouville numbers.

Using the explicit continued fraction expansion of e, one can show that e is not a Liouville number (although the partial quotients in its continued fraction expansion are unbounded). Kurt Mahler showed in 1953 that π is also not a Liouville number. It is conjectured that all infinite continued fractions with bounded terms, that have a "simple" structure, and that are not eventually periodic are transcendental[18] (in other words, algebraic irrational roots of at least third degree polynomials do not have apparent pattern in their continued fraction expansions, since eventually periodic continued fractions correspond to quadratic irrationals, see Hermite's problem).

Numbers proven to be transcendental[edit]

Numbers proven to be transcendental:

, the Gelfond–Schneider constant (or Hilbert number)
  • sin a, cos a, tan a, csc a, sec a, and cot a, and their hyperbolic counterparts, for any nonzero algebraic number a, expressed in radians (by the Lindemann–Weierstrass theorem).
  • The fixed point of the cosine function (also referred to as the Dottie number d) – the unique real solution to the equation cos x = x, where x is in radians (by the Lindemann–Weierstrass theorem).[19]
  • ln a if a is algebraic and not equal to 0 or 1, for any branch of the logarithm function (by the Lindemann–Weierstrass theorem), in particular: the universal parabolic constant.
  • logb a if a and b are positive integers not both powers of the same integer, and a is not equal to 1 (by the Gelfond–Schneider theorem).
  • Non-zero results of arcsin a, arccos a, arctan a, arccsc a, arcsec a, arccot a and their hyperbolic counterparts, for any algebraic number a (by the Lindemann–Weierstrass theorem).
  • The Bessel function of the first kind Jν(x), its first derivative, and the quotient are transcendental when ν is rational and x is algebraic and nonzero,[20] and all nonzero roots of Jν(x) and J'ν(x) are transcendental when ν is rational.[21]
  • W(a) if a is algebraic and nonzero, for any branch of the Lambert W Function (by the Lindemann–Weierstrass theorem), in particular: Ω the omega constant
  • W(r,a) if both a and the order r are algebraic such that , for any branch of the generalized Lambert W function.[22]
  • xs, the square super-root of any natural number is either an integer or transcendental (by the Gelfond–Schneider theorem)
  • ,[23] ,[24] and .[24] The numbers and are also known to be transcendental. The numbers and are also transcendental.[25]
  • The values of Euler beta function (in which a, b and are non-integer rational numbers).[26]
  • 0.64341054629 ... , Cahen's constant.[27]
  • .[28] In general, all numbers of the form are transcendental, where are algebraic for all and are non-zero algebraic for all (by the Baker's theorem).
  • The Champernowne constants, the irrational numbers formed by concatenating representations of all positive integers.[29]
  • Ω, Chaitin's constant (since it is a non-computable number).[30]
  • The supremum limit of the Specker sequences (since they are non-computable numbers).[31]
  • The so-called Fredholm constants, such as[11][32][b]
which also holds by replacing 10 with any algebraic number b > 1.[34]
  • , for rational number x such that .[28]
  • The values of the Rogers-Ramanujan continued fraction where is algebraic and .[35] The lemniscatic values of theta function (under the same conditions for ) are also transcendental.[36]
  • j(q) where is algebraic but not imaginary quadratic (i.e, the exceptional set of this function is the number field whose degree of extension over is 2).
  • The values of the infinite series with fast convergence rate as defined by Y. Gao and J. Gao, such as .[37]
  • The real constant in the definition of van der Corput's constant involving the Fresnel integrals.[38]
  • The real constant in the definition of Zolotarev-Schur constant involving the complete elliptic integral functions.[39]
  • Gauss's constant and the related lemniscate constant.[40]
  • Any number of the form (where , are polynomials in variables and , is algebraic and , is any integer greater than 1).[41]
  • Artificially constructed non-periodic numbers.[42]
  • The Robbins constant in three-dimensional line picking problem.[43]
  • The aforementioned Liouville constant for any algebraic b ∈ (0, 1).
  • The sum of reciprocals of exponential factorials.[28]
  • The Prouhet–Thue–Morse constant[44] and the related rabbit constant.[45]
  • The Komornik–Loreti constant.[46]
  • Any number for which the digits with respect to some fixed base form a Sturmian word.[47]
  • The paperfolding constant (also named as "Gaussian Liouville number").[48]
  • Constructed irrational numbers which are not simply normal in any base.[49]
  • For β > 1
where is the floor function.[50]
  • 3.300330000000000330033... and its reciprocal 0.30300000303..., two numbers with only two different decimal digits whose nonzero digit positions are given by the Moser–de Bruijn sequence and its double.[51]
  • The number , where Yα(x) and Jα(x) are Bessel functions and γ is the Euler–Mascheroni constant.[52][53]
  • Nesterenko proved in 1996 that and are algebraically independent.[25] This results in the transcendence of the Weierstrass constant[54] and the number .[55]

Possible transcendental numbers[edit]

Numbers which have yet to be proven to be either transcendental or algebraic:

  • Most sums, products, powers, etc. of the number π and the number e, e.g. , e + π, πe, π/e, ππ, ee, πe, π2, eπ2 are not known to be rational, algebraically irrational or transcendental. A notable exception is eπn (for any positive integer n) which has been proven transcendental.[56] It has been shown that both e + π and π/e do not satisfy any polynomial equation of degree and integer coefficients of average size 109.[57]
  • The Euler–Mascheroni constant γ: In 2010 M. Ram Murty and N. Saradha found an infinite list of numbers containing γ/4 such that all but at most one of them are transcendental.[58][59] In 2012 it was shown that at least one of γ and the Euler–Gompertz constant δ is transcendental.[60]
  • Apéry's constant ζ(3) (whose irrationality was proved by Apéry).
  • The reciprocal Fibonacci constant and reciprocal Lucas constant[61] (both of which have been proved to be irrational).
  • Catalan's constant, and the values of Dirichlet beta function at other even integers, β(4), β(6), ... (not even proven to be irrational).[62]
  • Khinchin's constant, also not proven to be irrational.
  • The Riemann zeta function at other odd positive integers, ζ(5), ζ(7), ... (not proven to be irrational).
  • The Feigenbaum constants δ and α, also not proven to be irrational.
  • Mills' constant and twin prime constant (also not proven to be irrational).
  • The cube super-root of any natural number is either an integer or irrational (by the Gelfond–Schneider theorem).[63] However, it is still unclear if the irrational numbers in the later case are all transcendental.[citation needed]
  • The second and later eigenvalues of the Gauss-Kuzmin-Wirsing operator, also not proven to be irrational.
  • The Copeland–Erdős constant, formed by concatenating the decimal representations of the prime numbers.
  • The relative density of regular prime numbers: in 1964, Siegel conjectured that its value is .
  • has not been proven to be irrational.[25]
  • Various constants whose value is not known with high precision, such as the Landau's constant and the Grothendieck constant.

Related conjectures:

Proofs for specific numbers[edit]

A proof that e is transcendental[edit]

The first proof that the base of the natural logarithms, e, is transcendental dates from 1873. We will now follow the strategy of David Hilbert (1862–1943) who gave a simplification of the original proof of Charles Hermite. The idea is the following:

Assume, for purpose of finding a contradiction, that e is algebraic. Then there exists a finite set of integer coefficients c0, c1, ..., cn satisfying the equation:

It is difficult to make use of the integer status of these coefficients when multiplied by a power of the irrational e, but we can absorb those powers into an integral which “mostly” will assume integer values. For a positive integer k, define the polynomial
and multiply both sides of the above equation by
to arrive at the equation:

By splitting respective domains of integration, this equation can be written in the form

where
Here P will turn out to be an integer, but more importantly it grows quickly with k.

Lemma 1[edit]

There are arbitrarily large k such that is a non-zero integer.

Proof. Recall the standard integral (case of the Gamma function)

valid for any natural number . More generally,

if then .

This would allow us to compute exactly, because any term of can be rewritten as

through a change of variables. Hence
That latter sum is a polynomial in with integer coefficients, i.e., it is a linear combination of powers with integer coefficients. Hence the number is a linear combination (with those same integer coefficients) of factorials ; in particular is an integer.

Smaller factorials divide larger factorials, so the smallest occurring in that linear combination will also divide the whole of . We get that from the lowest power term appearing with a nonzero coefficient in , but this smallest exponent is also the multiplicity of as a root of this polynomial. is chosen to have multiplicity of the root and multiplicity of the roots for , so that smallest exponent is for and for with . Therefore divides .

To establish the last claim in the lemma, that is nonzero, it is sufficient to prove that does not divide . To that end, let be any prime larger than and . We know from the above that divides each of for , so in particular all of those are divisible by . It comes down to the first term . We have (see falling and rising factorials)

and those higher degree terms all give rise to factorials or larger. Hence
That right hand side is a product of nonzero integer factors less than the prime , therefore that product is not divisible by , and the same holds for ; in particular cannot be zero.

Lemma 2[edit]

For sufficiently large k, .

Proof. Note that

where u(x), v(x) are continuous functions of x for all x, so are bounded on the interval [0, n]. That is, there are constants G, H > 0 such that

So each of those integrals composing Q is bounded, the worst case being

It is now possible to bound the sum Q as well:

where M is a constant not depending on k. It follows that

finishing the proof of this lemma.

Conclusion[edit]

Choosing a value of k satisfying both lemmas leads to a non-zero integer added to a vanishingly small quantity being equal to zero, is an impossibility. It follows that the original assumption, that e can satisfy a polynomial equation with integer coefficients, is also impossible; that is, e is transcendental.

The transcendence of π[edit]

A similar strategy, different from Lindemann's original approach, can be used to show that the number π is transcendental. Besides the gamma-function and some estimates as in the proof for e, facts about symmetric polynomials play a vital role in the proof.

For detailed information concerning the proofs of the transcendence of π and e, see the references and external links.

See also[edit]

Number systems
Complex
Real
Rational
Integer
Natural
Zero: 0
One: 1
Prime numbers
Composite numbers
Negative integers
Fraction
Finite decimal
Dyadic (finite binary)
Repeating decimal
Irrational
Algebraic irrational
Transcendental
Imaginary

Notes[edit]

  1. ^ Cantor's construction builds a one-to-one correspondence between the set of transcendental numbers and the set of real numbers. In this article, Cantor only applies his construction to the set of irrational numbers.[15]
  2. ^ The name 'Fredholm number' is misplaced: Kempner first proved this number is transcendental, and the note on page 403 states that Fredholm never studied this number.[33]

References[edit]

  1. ^ Pickover, Cliff. "The 15 most famous transcendental numbers". sprott.physics.wisc.edu. Retrieved 2020-01-23.
  2. ^ Shidlovskii, Andrei B. (June 2011). Transcendental Numbers. Walter de Gruyter. p. 1. ISBN 9783110889055.
  3. ^ a b Bunday, B. D.; Mulholland, H. (20 May 2014). Pure Mathematics for Advanced Level. Butterworth-Heinemann. ISBN 978-1-4831-0613-7. Retrieved 21 March 2021.
  4. ^ Baker, A. (1964). "On Mahler's classification of transcendental numbers". Acta Mathematica. 111: 97–120. doi:10.1007/bf02391010. S2CID 122023355.
  5. ^ Heuer, Nicolaus; Loeh, Clara (1 November 2019). "Transcendental simplicial volumes". arXiv:1911.06386 [math.GT].
  6. ^ "Real number". Encyclopædia Britannica. mathematics. Retrieved 2020-08-11.
  7. ^ "transcendental". Oxford English Dictionary. s.v.
  8. ^ Leibniz, Gerhardt & Pertz 1858, pp. 97–98; Bourbaki 1994, p. 74
  9. ^ Erdős & Dudley 1983
  10. ^ Lambert 1768
  11. ^ a b Kempner 1916
  12. ^ "Weisstein, Eric W. "Liouville's Constant", MathWorld".
  13. ^ Liouville 1851
  14. ^ Cantor 1874; Gray 1994
  15. ^ Cantor 1878, p. 254
  16. ^ Baker, Alan (1998). J.J. O'Connor and E.F. Robertson. www-history.mcs.st-andrews.ac.uk (biographies). The MacTutor History of Mathematics archive. St. Andrew's, Scotland: University of St. Andrew's.
  17. ^ Hardy 1979
  18. ^ Adamczewski & Bugeaud 2005
  19. ^ Weisstein, Eric W. "Dottie Number". Wolfram MathWorld. Wolfram Research, Inc. Retrieved 23 July 2016.
  20. ^ Siegel, Carl L. (2014). "Über einige Anwendungen diophantischer Approximationen: Abhandlungen der Preußischen Akademie der Wissenschaften. Physikalisch-mathematische Klasse 1929, Nr. 1". On Some Applications of Diophantine Approximations (in German). Scuola Normale Superiore. pp. 81–138. doi:10.1007/978-88-7642-520-2_2. ISBN 978-88-7642-520-2.
  21. ^ Lorch, Lee; Muldoon, Martin E. (1995). "Transcendentality of zeros of higher dereivatives of functions involving Bessel functions". International Journal of Mathematics and Mathematical Sciences. 18 (3): 551–560. doi:10.1155/S0161171295000706.
  22. ^ Mező, István; Baricz, Árpád (June 22, 2015). "On the generalization of the Lambert W function". arXiv:1408.3999 [math.CA].
  23. ^ le Lionnais 1979, p. 46 via Wolfram Mathworld, Transcendental Number
  24. ^ a b Chudnovsky 1984 via Wolfram Mathworld, Transcendental Number
  25. ^ a b c "Mathematical constants". Mathematics (general). Cambridge University Press. Retrieved 2022-09-22.
  26. ^ Waldschmidt, Michel (September 7, 2005). "Transcendence of Periods: The State of the Art" (PDF). webusers.imj-prg.fr.
  27. ^ Davison & Shallit 1991
  28. ^ a b c Weisstein, Eric W. "Transcendental Number". mathworld.wolfram.com. Retrieved 2023-08-09.
  29. ^ Mahler 1937; Mahler 1976, p. 12
  30. ^ Calude 2002, p. 239
  31. ^ Grue Simonsen, Jakob. "Specker Sequences Revisited" (PDF). hjemmesider.diku.dk.
  32. ^ Shallit 1996
  33. ^ Allouche & Shallit 2003, pp. 385, 403
  34. ^ Loxton 1988
  35. ^ Duverney, Daniel; Nishioka, Keiji; Nishioka, Kumiko; Shiokawa, Iekata (1997). "Transcendence of Rogers-Ramanujan continued fraction and reciprocal sums of Fibonacci numbers". Proceedings of the Japan Academy, Series A, Mathematical Sciences. 73 (7): 140–142. doi:10.3792/pjaa.73.140. ISSN 0386-2194.
  36. ^ Bertrand, Daniel (1997). "Theta functions and transcendence". The Ramanujan Journal. 1 (4): 339–350. doi:10.1023/A:1009749608672. S2CID 118628723.
  37. ^ "A140654 - OEIS". oeis.org. Retrieved 2023-08-12.
  38. ^ Weisstein, Eric W. "van der Corput's Constant". mathworld.wolfram.com. Retrieved 2023-08-10.
  39. ^ Weisstein, Eric W. "Zolotarev-Schur Constant". mathworld.wolfram.com. Retrieved 2023-08-12.
  40. ^ Todd, John (1975). "The lemniscate constants". Communications of the ACM. 18: 14–19. doi:10.1145/360569.360580. S2CID 85873.
  41. ^ Kurosawa, Takeshi (2007-03-01). "Transcendence of certain series involving binary linear recurrences". Journal of Number Theory. 123 (1): 35–58. doi:10.1016/j.jnt.2006.05.019. ISSN 0022-314X.
  42. ^ Yoshinaga, Masahiko (2008-05-03). "Periods and elementary real numbers". arXiv:0805.0349 [math.AG].
  43. ^ Steven R. Finch (2003). Mathematical Constants. Cambridge University Press. p. 479. ISBN 978-3-540-67695-9. Schmutz.
  44. ^ Mahler 1929; Allouche & Shallit 2003, p. 387
  45. ^ Weisstein, Eric W. "Rabbit Constant". mathworld.wolfram.com. Retrieved 2023-08-09.
  46. ^ Allouche, Jean-Paul; Cosnard, Michel (2000), "The Komornik–Loreti constant is transcendental", American Mathematical Monthly, 107 (5): 448–449, doi:10.2307/2695302, JSTOR 2695302, MR 1763399
  47. ^ Pytheas Fogg 2002
  48. ^ "A143347 - OEIS". oeis.org. Retrieved 2023-08-09.
  49. ^ Bugeaud 2012, p. 113.
  50. ^ Adamczewski, Boris (March 2013). "The Many Faces of the Kempner Number". arXiv:1303.1685 [math.NT].
  51. ^ Blanchard & Mendès France 1982
  52. ^ Mahler, Kurt; Mordell, Louis Joel (1968-06-04). "Applications of a theorem by A. B. Shidlovski". Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences. 305 (1481): 149–173. Bibcode:1968RSPSA.305..149M. doi:10.1098/rspa.1968.0111. S2CID 123486171.
  53. ^ Lagarias, Jeffrey C. (2013-07-19). "Euler's constant: Euler's work and modern developments". Bulletin of the American Mathematical Society. 50 (4): 527–628. arXiv:1303.1856. doi:10.1090/S0273-0979-2013-01423-X. ISSN 0273-0979.
  54. ^ Weisstein, Eric W. "Weierstrass Constant". mathworld.wolfram.com. Retrieved 2023-08-12.
  55. ^ Elsner, Carsten; Shimomura, Shun; Shiokawa, Iekata (2012-09-01). "Algebraic independence of certain numbers related to modular functions". Functiones et Approximatio Commentarii Mathematici. 47 (1). doi:10.7169/facm/2012.47.1.10. ISSN 0208-6573.
  56. ^ Weisstein, Eric W. "Irrational Number". MathWorld.
  57. ^ Weisstein, Eric W. "e". mathworld.wolfram.com. Retrieved 2023-08-12.
  58. ^ Murty, M. Ram; Saradha, N. (2010-12-01). "Euler–Lehmer constants and a conjecture of Erdös". Journal of Number Theory. 130 (12): 2671–2682. doi:10.1016/j.jnt.2010.07.004. ISSN 0022-314X.
  59. ^ Murty, M. Ram; Zaytseva, Anastasia (2013-01-01). "Transcendence of generalized Euler constants". The American Mathematical Monthly. 120 (1): 48–54. doi:10.4169/amer.math.monthly.120.01.048. ISSN 0002-9890. S2CID 20495981.
  60. ^ Rivoal, Tanguy (2012). "On the arithmetic nature of the values of the gamma function, Euler's constant, and Gompertz's constant". Michigan Mathematical Journal. 61 (2): 239–254. doi:10.1307/mmj/1339011525. ISSN 0026-2285.
  61. ^ "A093540 - OEIS". oeis.org. Retrieved 2023-08-12.
  62. ^ Rivoal, T.; Zudilin, W. (2003-08-01). "Diophantine properties of numbers related to Catalan's constant". Mathematische Annalen. 326 (4): 705–721. doi:10.1007/s00208-003-0420-2. hdl:1959.13/803688. ISSN 1432-1807. S2CID 59328860.
  63. ^ Marshall, J. Ash; Tan, Yiren (March 2012). "A rational number of the form aa with a irrational" (PDF).

Sources[edit]

External links[edit]