fbpx
Wikipedia

Irrational number

In mathematics, the irrational numbers (in- + rational) are all the real numbers that are not rational numbers. That is, irrational numbers cannot be expressed as the ratio of two integers. When the ratio of lengths of two line segments is an irrational number, the line segments are also described as being incommensurable, meaning that they share no "measure" in common, that is, there is no length ("the measure"), no matter how short, that could be used to express the lengths of both of the two given segments as integer multiples of itself.

The number 2 is irrational.

Among irrational numbers are the ratio π of a circle's circumference to its diameter, Euler's number e, the golden ratio φ, and the square root of two.[1] In fact, all square roots of natural numbers, other than of perfect squares, are irrational.[2]

Like all real numbers, irrational numbers can be expressed in positional notation, notably as a decimal number. In the case of irrational numbers, the decimal expansion does not terminate, nor end with a repeating sequence. For example, the decimal representation of π starts with 3.14159, but no finite number of digits can represent π exactly, nor does it repeat. Conversely, a decimal expansion that terminates or repeats must be a rational number. These are provable properties of rational numbers and positional number systems and are not used as definitions in mathematics.

Irrational numbers can also be expressed as non-terminating continued fractions (which in some cases are periodic), and in many other ways.

As a consequence of Cantor's proof that the real numbers are uncountable and the rationals countable, it follows that almost all real numbers are irrational.[3]

History edit

 
Set of real numbers (R), which include the rationals (Q), which include the integers (Z), which include the natural numbers (N). The real numbers also include the irrationals (R\Q).

Ancient Greece edit

The first proof of the existence of irrational numbers is usually attributed to a Pythagorean (possibly Hippasus of Metapontum),[4] who probably discovered them while identifying sides of the pentagram.[5] The Pythagorean method would have claimed that there must be some sufficiently small, indivisible unit that could fit evenly into one of these lengths as well as the other. Hippasus in the 5th century BC, however, was able to deduce that there was no common unit of measure, and that the assertion of such an existence was a contradiction. He did this by demonstrating that if the hypotenuse of an isosceles right triangle was indeed commensurable with a leg, then one of those lengths measured in that unit of measure must be both odd and even, which is impossible. His reasoning is as follows:

  • Start with an isosceles right triangle with side lengths of integers a, b, and c. The ratio of the hypotenuse to a leg is represented by c:b.
  • Assume a, b, and c are in the smallest possible terms (i.e. they have no common factors).
  • By the Pythagorean theorem: c2 = a2+b2 = b2+b2 = 2b2. (Since the triangle is isosceles, a = b).
  • Since c2 = 2b2, c2 is divisible by 2, and therefore even.
  • Since c2 is even, c must be even.
  • Since c is even, dividing c by 2 yields an integer. Let y be this integer (c = 2y).
  • Squaring both sides of c = 2y yields c2 = (2y)2, or c2 = 4y2.
  • Substituting 4y2 for c2 in the first equation (c2 = 2b2) gives us 4y2= 2b2.
  • Dividing by 2 yields 2y2 = b2.
  • Since y is an integer, and 2y2 = b2, b2 is divisible by 2, and therefore even.
  • Since b2 is even, b must be even.
  • We have just shown that both b and c must be even. Hence they have a common factor of 2. However, this contradicts the assumption that they have no common factors. This contradiction proves that c and b cannot both be integers and thus the existence of a number that cannot be expressed as a ratio of two integers.[6]

Greek mathematicians termed this ratio of incommensurable magnitudes alogos, or inexpressible. Hippasus, however, was not lauded for his efforts: according to one legend, he made his discovery while out at sea, and was subsequently thrown overboard by his fellow Pythagoreans 'for having produced an element in the universe which denied the... doctrine that all phenomena in the universe can be reduced to whole numbers and their ratios.'[7] Another legend states that Hippasus was merely exiled for this revelation. Whatever the consequence to Hippasus himself, his discovery posed a very serious problem to Pythagorean mathematics, since it shattered the assumption that numbers and geometry were inseparable; a foundation of their theory.

The discovery of incommensurable ratios was indicative of another problem facing the Greeks: the relation of the discrete to the continuous. This was brought to light by Zeno of Elea, who questioned the conception that quantities are discrete and composed of a finite number of units of a given size. Past Greek conceptions dictated that they necessarily must be, for "whole numbers represent discrete objects, and a commensurable ratio represents a relation between two collections of discrete objects",[8] but Zeno found that in fact "[quantities] in general are not discrete collections of units; this is why ratios of incommensurable [quantities] appear... .[Q]uantities are, in other words, continuous".[8] What this means is that contrary to the popular conception of the time, there cannot be an indivisible, smallest unit of measure for any quantity. In fact, these divisions of quantity must necessarily be infinite. For example, consider a line segment: this segment can be split in half, that half split in half, the half of the half in half, and so on. This process can continue infinitely, for there is always another half to be split. The more times the segment is halved, the closer the unit of measure comes to zero, but it never reaches exactly zero. This is just what Zeno sought to prove. He sought to prove this by formulating four paradoxes, which demonstrated the contradictions inherent in the mathematical thought of the time. While Zeno's paradoxes accurately demonstrated the deficiencies of contemporary mathematical conceptions, they were not regarded as proof of the alternative. In the minds of the Greeks, disproving the validity of one view did not necessarily prove the validity of another, and therefore, further investigation had to occur.

The next step was taken by Eudoxus of Cnidus, who formalized a new theory of proportion that took into account commensurable as well as incommensurable quantities. Central to his idea was the distinction between magnitude and number. A magnitude "...was not a number but stood for entities such as line segments, angles, areas, volumes, and time which could vary, as we would say, continuously. Magnitudes were opposed to numbers, which jumped from one value to another, as from 4 to 5".[9] Numbers are composed of some smallest, indivisible unit, whereas magnitudes are infinitely reducible. Because no quantitative values were assigned to magnitudes, Eudoxus was then able to account for both commensurable and incommensurable ratios by defining a ratio in terms of its magnitude, and proportion as an equality between two ratios. By taking quantitative values (numbers) out of the equation, he avoided the trap of having to express an irrational number as a number. "Eudoxus' theory enabled the Greek mathematicians to make tremendous progress in geometry by supplying the necessary logical foundation for incommensurable ratios".[10] This incommensurability is dealt with in Euclid's Elements, Book X, Proposition 9. It was not until Eudoxus developed a theory of proportion that took into account irrational as well as rational ratios that a strong mathematical foundation of irrational numbers was created.[11]

As a result of the distinction between number and magnitude, geometry became the only method that could take into account incommensurable ratios. Because previous numerical foundations were still incompatible with the concept of incommensurability, Greek focus shifted away from numerical conceptions such as algebra and focused almost exclusively on geometry. In fact, in many cases, algebraic conceptions were reformulated into geometric terms. This may account for why we still conceive of x2 and x3 as x squared and x cubed instead of x to the second power and x to the third power. Also crucial to Zeno's work with incommensurable magnitudes was the fundamental focus on deductive reasoning that resulted from the foundational shattering of earlier Greek mathematics. The realization that some basic conception within the existing theory was at odds with reality necessitated a complete and thorough investigation of the axioms and assumptions that underlie that theory. Out of this necessity, Eudoxus developed his method of exhaustion, a kind of reductio ad absurdum that "...established the deductive organization on the basis of explicit axioms..." as well as "...reinforced the earlier decision to rely on deductive reasoning for proof".[12] This method of exhaustion is the first step in the creation of calculus.

Theodorus of Cyrene proved the irrationality of the surds of whole numbers up to 17, but stopped there probably because the algebra he used could not be applied to the square root of 17.[13]

India edit

Geometrical and mathematical problems involving irrational numbers such as square roots were addressed very early during the Vedic period in India. There are references to such calculations in the Samhitas, Brahmanas, and the Shulba Sutras (800 BC or earlier).[14]

It is suggested that the concept of irrationality was implicitly accepted by Indian mathematicians since the 7th century BC, when Manava (c. 750 – 690 BC) believed that the square roots of numbers such as 2 and 61 could not be exactly determined.[15] Historian Carl Benjamin Boyer, however, writes that "such claims are not well substantiated and unlikely to be true".[16]

Later, in their treatises, Indian mathematicians wrote on the arithmetic of surds including addition, subtraction, multiplication, rationalization, as well as separation and extraction of square roots.[17]

Mathematicians like Brahmagupta (in 628 AD) and Bhāskara I (in 629 AD) made contributions in this area as did other mathematicians who followed. In the 12th century Bhāskara II evaluated some of these formulas and critiqued them, identifying their limitations.

During the 14th to 16th centuries, Madhava of Sangamagrama and the Kerala school of astronomy and mathematics discovered the infinite series for several irrational numbers such as π and certain irrational values of trigonometric functions. Jyeṣṭhadeva provided proofs for these infinite series in the Yuktibhāṣā.[18]

Islamic World edit

In the Middle Ages, the development of algebra by Muslim mathematicians allowed irrational numbers to be treated as algebraic objects.[19] Middle Eastern mathematicians also merged the concepts of "number" and "magnitude" into a more general idea of real numbers, criticized Euclid's idea of ratios, developed the theory of composite ratios, and extended the concept of number to ratios of continuous magnitude.[20] In his commentary on Book 10 of the Elements, the Persian mathematician Al-Mahani (d. 874/884) examined and classified quadratic irrationals and cubic irrationals. He provided definitions for rational and irrational magnitudes, which he treated as irrational numbers. He dealt with them freely but explains them in geometric terms as follows:[20]

"It will be a rational (magnitude) when we, for instance, say 10, 12, 3%, 6%, etc., because its value is pronounced and expressed quantitatively. What is not rational is irrational and it is impossible to pronounce and represent its value quantitatively. For example: the roots of numbers such as 10, 15, 20 which are not squares, the sides of numbers which are not cubes etc."

In contrast to Euclid's concept of magnitudes as lines, Al-Mahani considered integers and fractions as rational magnitudes, and square roots and cube roots as irrational magnitudes. He also introduced an arithmetical approach to the concept of irrationality, as he attributes the following to irrational magnitudes:[20]

"their sums or differences, or results of their addition to a rational magnitude, or results of subtracting a magnitude of this kind from an irrational one, or of a rational magnitude from it."

The Egyptian mathematician Abū Kāmil Shujā ibn Aslam (c. 850 – 930) was the first to accept irrational numbers as solutions to quadratic equations or as coefficients in an equation in the form of square roots and fourth roots.[21] In the 10th century, the Iraqi mathematician Al-Hashimi provided general proofs (rather than geometric demonstrations) for irrational numbers, as he considered multiplication, division, and other arithmetical functions.[20]

Many of these concepts were eventually accepted by European mathematicians sometime after the Latin translations of the 12th century. Al-Hassār, a Moroccan mathematician from Fez specializing in Islamic inheritance jurisprudence during the 12th century, first mentions the use of a fractional bar, where numerators and denominators are separated by a horizontal bar. In his discussion he writes, "..., for example, if you are told to write three-fifths and a third of a fifth, write thus,  ."[22] This same fractional notation appears soon after in the work of Leonardo Fibonacci in the 13th century.[23]

Modern period edit

The 17th century saw imaginary numbers become a powerful tool in the hands of Abraham de Moivre, and especially of Leonhard Euler. The completion of the theory of complex numbers in the 19th century entailed the differentiation of irrationals into algebraic and transcendental numbers, the proof of the existence of transcendental numbers, and the resurgence of the scientific study of the theory of irrationals, largely ignored since Euclid. The year 1872 saw the publication of the theories of Karl Weierstrass (by his pupil Ernst Kossak), Eduard Heine (Crelle's Journal, 74), Georg Cantor (Annalen, 5), and Richard Dedekind. Méray had taken in 1869 the same point of departure as Heine, but the theory is generally referred to the year 1872. Weierstrass's method has been completely set forth by Salvatore Pincherle in 1880,[24] and Dedekind's has received additional prominence through the author's later work (1888) and the endorsement by Paul Tannery (1894). Weierstrass, Cantor, and Heine base their theories on infinite series, while Dedekind founds his on the idea of a cut (Schnitt) in the system of all rational numbers, separating them into two groups having certain characteristic properties. The subject has received later contributions at the hands of Weierstrass, Leopold Kronecker (Crelle, 101), and Charles Méray.

Continued fractions, closely related to irrational numbers (and due to Cataldi, 1613), received attention at the hands of Euler, and at the opening of the 19th century were brought into prominence through the writings of Joseph-Louis Lagrange. Dirichlet also added to the general theory, as have numerous contributors to the applications of the subject.

Johann Heinrich Lambert proved (1761) that π cannot be rational, and that en is irrational if n is rational (unless n = 0).[25] While Lambert's proof is often called incomplete, modern assessments support it as satisfactory, and in fact for its time it is unusually rigorous. Adrien-Marie Legendre (1794), after introducing the Bessel–Clifford function, provided a proof to show that π2 is irrational, whence it follows immediately that π is irrational also. The existence of transcendental numbers was first established by Liouville (1844, 1851). Later, Georg Cantor (1873) proved their existence by a different method, which showed that every interval in the reals contains transcendental numbers. Charles Hermite (1873) first proved e transcendental, and Ferdinand von Lindemann (1882), starting from Hermite's conclusions, showed the same for π. Lindemann's proof was much simplified by Weierstrass (1885), still further by David Hilbert (1893), and was finally made elementary by Adolf Hurwitz[citation needed] and Paul Gordan.[26]

Examples edit

Square roots edit

The square root of 2 was likely the first number proved irrational.[27] The golden ratio is another famous quadratic irrational number. The square roots of all natural numbers that are not perfect squares are irrational and a proof may be found in quadratic irrationals.

General roots edit

The proof above[clarification needed] for the square root of two can be generalized using the fundamental theorem of arithmetic. This asserts that every integer has a unique factorization into primes. Using it we can show that if a rational number is not an integer then no integral power of it can be an integer, as in lowest terms there must be a prime in the denominator that does not divide into the numerator whatever power each is raised to. Therefore, if an integer is not an exact kth power of another integer, then that first integer's kth root is irrational.

Logarithms edit

Perhaps the numbers most easy to prove irrational are certain logarithms. Here is a proof by contradiction that log2 3 is irrational (log2 3 ≈ 1.58 > 0).

Assume log2 3 is rational. For some positive integers m and n, we have

 

It follows that

 
 
 

The number 2 raised to any positive integer power must be even (because it is divisible by 2) and the number 3 raised to any positive integer power must be odd (since none of its prime factors will be 2). Clearly, an integer cannot be both odd and even at the same time: we have a contradiction. The only assumption we made was that log2 3 is rational (and so expressible as a quotient of integers m/n with n ≠ 0). The contradiction means that this assumption must be false, i.e. log2 3 is irrational, and can never be expressed as a quotient of integers m/n with n ≠ 0.

Cases such as log10 2 can be treated similarly.

Types edit

An irrational number may be algebraic, that is a real root of a polynomial with integer coefficients. Those that are not algebraic are transcendental.

Algebraic edit

The real algebraic numbers are the real solutions of polynomial equations

 

where the coefficients   are integers and  . An example of an irrational algebraic number is x0 = (21/2 + 1)1/3. It is clearly algebraic since it is the root of an integer polynomial, (x3 − 1)2 = 2, which is equivalent to x6 − 2x3 − 1 = 0. This polynomial has no rational roots, since the rational root theorem shows that the only possibilities are ±1, but x0is greater than 1. So x0 is an irrational algebraic number. There are countably many algebraic numbers, since there are countably many integer polynomials.

Transcendental edit

So almost all irrational numbers are transcendental. Examples are e r and π r, which are transcendental for all nonzero rational r.

Because the algebraic numbers form a subfield of the real numbers, many irrational real numbers can be constructed by combining transcendental and algebraic numbers. For example, 3π + 2, π + 2 and e3 are irrational (and even transcendental).

Decimal expansions edit

The decimal expansion of an irrational number never repeats (meaning the decimal expansion does not repeat the same number or sequence of numbers) or terminates (this means there is not a finite number of nonzero digits), unlike any rational number. The same is true for binary, octal or hexadecimal expansions, and in general for expansions in every positional notation with natural bases.

To show this, suppose we divide integers n by m (where m is nonzero). When long division is applied to the division of n by m, there can never be a remainder greater than or equal to m. If 0 appears as a remainder, the decimal expansion terminates. If 0 never occurs, then the algorithm can run at most m − 1 steps without using any remainder more than once. After that, a remainder must recur, and then the decimal expansion repeats.

Conversely, suppose we are faced with a repeating decimal, we can prove that it is a fraction of two integers. For example, consider:

 

Here the repetend is 162 and the length of the repetend is 3. First, we multiply by an appropriate power of 10 to move the decimal point to the right so that it is just in front of a repetend. In this example we would multiply by 10 to obtain:

 

Now we multiply this equation by 10r where r is the length of the repetend. This has the effect of moving the decimal point to be in front of the "next" repetend. In our example, multiply by 103:

 

The result of the two multiplications gives two different expressions with exactly the same "decimal portion", that is, the tail end of 10,000A matches the tail end of 10A exactly. Here, both 10,000A and 10A have .162162162... after the decimal point.

Therefore, when we subtract the 10A equation from the 10,000A equation, the tail end of 10A cancels out the tail end of 10,000A leaving us with:

 

Then

 

is a ratio of integers and therefore a rational number.

Irrational powers edit

Dov Jarden gave a simple non-constructive proof that there exist two irrational numbers a and b, such that ab is rational:[28][29]

Consider 22; if this is rational, then take a = b = 2. Otherwise, take a to be the irrational number 22 and b = 2. Then ab = (22)2 = 22·2 = 22 = 2, which is rational.

Although the above argument does not decide between the two cases, the Gelfond–Schneider theorem shows that 22 is transcendental, hence irrational. This theorem states that if a and b are both algebraic numbers, and a is not equal to 0 or 1, and b is not a rational number, then any value of ab is a transcendental number (there can be more than one value if complex number exponentiation is used).

An example that provides a simple constructive proof is[30]

 

The base of the left side is irrational and the right side is rational, so one must prove that the exponent on the left side,  , is irrational. This is so because, by the formula relating logarithms with different bases,

 

which we can assume, for the sake of establishing a contradiction, equals a ratio m/n of positive integers. Then   hence   hence   hence  , which is a contradictory pair of prime factorizations and hence violates the fundamental theorem of arithmetic (unique prime factorization).

A stronger result is the following:[31] Every rational number in the interval   can be written either as aa for some irrational number a or as nn for some natural number n. Similarly,[31] every positive rational number can be written either as   for some irrational number a or as   for some natural number n.

Open questions edit

It is not known if   (or  ) is irrational. In fact, there is no pair of non-zero integers   for which it is known whether   is irrational. Moreover, it is not known if the set   is algebraically independent over  .

It is not known if   Catalan's constant, or the Euler–Mascheroni constant   are irrational.[32] It is not known if either of the tetrations   or   is rational for some integer  [citation needed]

In constructive mathematics edit

In constructive mathematics, excluded middle is not valid, so it is not true that every real number is rational or irrational. Thus, the notion of an irrational number bifurcates into multiple distinct notions. One could take the traditional definition of an irrational number as a real number that is not rational.[33] However, there is a second definition of an irrational number used in constructive mathematics, that a real number   is an irrational number if it is apart from every rational number, or equivalently, if the distance   between   and every rational number   is positive. This definition is stronger than the traditional definition of an irrational number. This second definition is used in Errett Bishop's proof that the square root of 2 is irrational.[34]

Set of all irrationals edit

Since the reals form an uncountable set, of which the rationals are a countable subset, the complementary set of irrationals is uncountable.

Under the usual (Euclidean) distance function  , the real numbers are a metric space and hence also a topological space. Restricting the Euclidean distance function gives the irrationals the structure of a metric space. Since the subspace of irrationals is not closed, the induced metric is not complete. Being a G-delta set—i.e., a countable intersection of open subsets—in a complete metric space, the space of irrationals is completely metrizable: that is, there is a metric on the irrationals inducing the same topology as the restriction of the Euclidean metric, but with respect to which the irrationals are complete. One can see this without knowing the aforementioned fact about G-delta sets: the continued fraction expansion of an irrational number defines a homeomorphism from the space of irrationals to the space of all sequences of positive integers, which is easily seen to be completely metrizable.

Furthermore, the set of all irrationals is a disconnected metrizable space. In fact, the irrationals equipped with the subspace topology have a basis of clopen groups so the space is zero-dimensional.

See also edit

Number systems
Complex  
Real  
Rational  
Integer  
Natural  
Negative integers
Imaginary

References edit

  1. ^ The 15 Most Famous Transcendental Numbers. by Clifford A. Pickover. URL retrieved 24 October 2007.
  2. ^ Jackson, Terence (2011-07-01). "95.42 Irrational square roots of natural numbers — a geometrical approach". The Mathematical Gazette. 95 (533): 327–330. doi:10.1017/S0025557200003193. ISSN 0025-5572. S2CID 123995083.
  3. ^ Cantor, Georg (1955) [1915]. Philip Jourdain (ed.). Contributions to the Founding of the Theory of Transfinite Numbers. New York: Dover. ISBN 978-0-486-60045-1.
  4. ^ Kurt Von Fritz (1945). "The Discovery of Incommensurability by Hippasus of Metapontum". Annals of Mathematics. 46 (2): 242–264. doi:10.2307/1969021. JSTOR 1969021. S2CID 126296119.
  5. ^ James R. Choike (1980). "The Pentagram and the Discovery of an Irrational Number". The Two-Year College Mathematics Journal. 11 (5): 312–316. doi:10.2307/3026893. JSTOR 3026893. S2CID 115390951.
  6. ^ Kline, M. (1990). Mathematical Thought from Ancient to Modern Times, Vol. 1. New York: Oxford University Press (original work published 1972), p. 33.
  7. ^ Kline 1990, p. 32.
  8. ^ a b Kline 1990, p. 34.
  9. ^ Kline 1990, p. 48.
  10. ^ Kline 1990, p. 49.
  11. ^ Charles H. Edwards (1982). The historical development of the calculus. Springer.
  12. ^ Kline 1990, p. 50.
  13. ^ Robert L. McCabe (1976). "Theodorus' Irrationality Proofs". Mathematics Magazine. 49 (4): 201–203. doi:10.1080/0025570X.1976.11976579. JSTOR 2690123. S2CID 124565880..
  14. ^ Bag, Amulya Kumar (1990). "Ritual Geometry in India and its Parallelism in other Culture Areas". Indian Journal of History of Science. 25.
  15. ^ T. K. Puttaswamy, "The Accomplishments of Ancient Indian Mathematicians", pp. 411–2, in Selin, Helaine; D'Ambrosio, Ubiratan, eds. (2000). Mathematics Across Cultures: The History of Non-western Mathematics. Springer. ISBN 1-4020-0260-2..
  16. ^ Boyer (1991). "China and India". A History of Mathematics (2nd ed.). Wiley. p. 208. ISBN 0471093742. OCLC 414892. It has been claimed also that the first recognition of incommensurables appears in India during the Sulbasutra period, but such claims are not well substantiated. The case for early Hindu awareness of incommensurable magnitudes is rendered most unlikely by the lack of evidence that Indian mathematicians of that period had come to grips with fundamental concepts.
  17. ^ Datta, Bibhutibhusan; Singh, Awadhesh Narayan (1993). (PDF). Indian Journal of History of Science. 28 (3): 253–264. Archived from the original (PDF) on 2018-10-03. Retrieved 18 September 2018.
  18. ^ Katz, V. J. (1995). "Ideas of Calculus in Islam and India". Mathematics Magazine. 63 (3): 163–174. doi:10.2307/2691411. JSTOR 2691411.
  19. ^ O'Connor, John J.; Robertson, Edmund F. (1999). "Arabic mathematics: forgotten brilliance?". MacTutor History of Mathematics Archive. University of St Andrews..
  20. ^ a b c d Matvievskaya, Galina (1987). "The theory of quadratic irrationals in medieval Oriental mathematics". Annals of the New York Academy of Sciences. 500 (1): 253–277. Bibcode:1987NYASA.500..253M. doi:10.1111/j.1749-6632.1987.tb37206.x. S2CID 121416910. See in particular pp. 254 & 259–260.
  21. ^ Jacques Sesiano, "Islamic mathematics", p. 148, in Selin, Helaine; D'Ambrosio, Ubiratan (2000). Mathematics Across Cultures: The History of Non-western Mathematics. Springer. ISBN 1-4020-0260-2..
  22. ^ Cajori, Florian (1928). A History of Mathematical Notations (Vol.1). La Salle, Illinois: The Open Court Publishing Company. pg. 269.
  23. ^ (Cajori 1928, pg.89)
  24. ^ Salvatore Pincherle (1880). "Saggio di una introduzione alla teoria delle funzioni analitiche secondo i principii del prof. C. Weierstrass". Giornale di Matematiche: 178–254, 317–320.
  25. ^ Lambert, J. H. (1761). "Mémoire sur quelques propriétés remarquables des quantités transcendentes, circulaires et logarithmiques" (PDF). Mémoires de l'Académie royale des sciences de Berlin (in French): 265–322. (PDF) from the original on 2016-04-28.
  26. ^ Gordan, Paul (1893). "Transcendenz von e und π". Mathematische Annalen. 43 (2–3). Teubner: 222–224. doi:10.1007/bf01443647. S2CID 123203471.
  27. ^ Fowler, David H. (2001), "The story of the discovery of incommensurability, revisited", Neusis (10): 45–61, MR 1891736
  28. ^ Jarden, Dov (1953). "Curiosa No. 339: A simple proof that a power of an irrational number to an irrational exponent may be rational". Scripta Mathematica. 19: 229. copy
  29. ^ George, Alexander; Velleman, Daniel J. (2002). Philosophies of mathematics (PDF). Blackwell. pp. 3–4. ISBN 0-631-19544-0.
  30. ^ Lord, Nick, "Maths bite: irrational powers of irrational numbers can be rational", Mathematical Gazette 92, November 2008, p. 534.
  31. ^ a b Marshall, Ash J., and Tan, Yiren, "A rational number of the form aa with a irrational", Mathematical Gazette 96, March 2012, pp. 106-109.
  32. ^ Albert, John. "Some unsolved problems in number theory" (PDF). Department of Mathematics, University of Oklahoma. (Senior Mathematics Seminar, Spring 2008 course)
  33. ^ Mark Bridger (2007). Real Analysis: A Constructive Approach through Interval Arithmetic. John Wiley & Sons. ISBN 978-1-470-45144-8.
  34. ^ Errett Bishop; Douglas Bridges (1985). Constructive Analysis. Springer. ISBN 0-387-15066-8.

Further reading edit

  • Adrien-Marie Legendre, Éléments de Géometrie, Note IV, (1802), Paris
  • Rolf Wallisser, "On Lambert's proof of the irrationality of π", in Algebraic Number Theory and Diophantine Analysis, Franz Halter-Koch and Robert F. Tichy, (2000), Walter de Gruyter

External links edit

  • Zeno's Paradoxes and Incommensurability 2016-05-13 at the Wayback Machine (n.d.). Retrieved April 1, 2008

irrational, number, mathematics, irrational, numbers, rational, real, numbers, that, rational, numbers, that, irrational, numbers, cannot, expressed, ratio, integers, when, ratio, lengths, line, segments, irrational, number, line, segments, also, described, be. In mathematics the irrational numbers in rational are all the real numbers that are not rational numbers That is irrational numbers cannot be expressed as the ratio of two integers When the ratio of lengths of two line segments is an irrational number the line segments are also described as being incommensurable meaning that they share no measure in common that is there is no length the measure no matter how short that could be used to express the lengths of both of the two given segments as integer multiples of itself The number 2 is irrational Among irrational numbers are the ratio p of a circle s circumference to its diameter Euler s number e the golden ratio f and the square root of two 1 In fact all square roots of natural numbers other than of perfect squares are irrational 2 Like all real numbers irrational numbers can be expressed in positional notation notably as a decimal number In the case of irrational numbers the decimal expansion does not terminate nor end with a repeating sequence For example the decimal representation of p starts with 3 14159 but no finite number of digits can represent p exactly nor does it repeat Conversely a decimal expansion that terminates or repeats must be a rational number These are provable properties of rational numbers and positional number systems and are not used as definitions in mathematics Irrational numbers can also be expressed as non terminating continued fractions which in some cases are periodic and in many other ways As a consequence of Cantor s proof that the real numbers are uncountable and the rationals countable it follows that almost all real numbers are irrational 3 Contents 1 History 1 1 Ancient Greece 1 2 India 1 3 Islamic World 1 4 Modern period 2 Examples 2 1 Square roots 2 2 General roots 2 3 Logarithms 3 Types 3 1 Algebraic 3 2 Transcendental 4 Decimal expansions 5 Irrational powers 6 Open questions 7 In constructive mathematics 8 Set of all irrationals 9 See also 10 References 11 Further reading 12 External linksHistory edit nbsp Set of real numbers R which include the rationals Q which include the integers Z which include the natural numbers N The real numbers also include the irrationals R Q Ancient Greece edit The first proof of the existence of irrational numbers is usually attributed to a Pythagorean possibly Hippasus of Metapontum 4 who probably discovered them while identifying sides of the pentagram 5 The Pythagorean method would have claimed that there must be some sufficiently small indivisible unit that could fit evenly into one of these lengths as well as the other Hippasus in the 5th century BC however was able to deduce that there was no common unit of measure and that the assertion of such an existence was a contradiction He did this by demonstrating that if the hypotenuse of an isosceles right triangle was indeed commensurable with a leg then one of those lengths measured in that unit of measure must be both odd and even which is impossible His reasoning is as follows Start with an isosceles right triangle with side lengths of integers a b and c The ratio of the hypotenuse to a leg is represented by c b Assume a b and c are in the smallest possible terms i e they have no common factors By the Pythagorean theorem c2 a2 b2 b2 b2 2b2 Since the triangle is isosceles a b Since c2 2b2 c2 is divisible by 2 and therefore even Since c2 is even c must be even Since c is even dividing c by 2 yields an integer Let y be this integer c 2y Squaring both sides of c 2y yields c2 2y 2 or c2 4y2 Substituting 4y2 for c2 in the first equation c2 2b2 gives us 4y2 2b2 Dividing by 2 yields 2y2 b2 Since y is an integer and 2y2 b2 b2 is divisible by 2 and therefore even Since b2 is even b must be even We have just shown that both b and c must be even Hence they have a common factor of 2 However this contradicts the assumption that they have no common factors This contradiction proves that c and b cannot both be integers and thus the existence of a number that cannot be expressed as a ratio of two integers 6 Greek mathematicians termed this ratio of incommensurable magnitudes alogos or inexpressible Hippasus however was not lauded for his efforts according to one legend he made his discovery while out at sea and was subsequently thrown overboard by his fellow Pythagoreans for having produced an element in the universe which denied the doctrine that all phenomena in the universe can be reduced to whole numbers and their ratios 7 Another legend states that Hippasus was merely exiled for this revelation Whatever the consequence to Hippasus himself his discovery posed a very serious problem to Pythagorean mathematics since it shattered the assumption that numbers and geometry were inseparable a foundation of their theory The discovery of incommensurable ratios was indicative of another problem facing the Greeks the relation of the discrete to the continuous This was brought to light by Zeno of Elea who questioned the conception that quantities are discrete and composed of a finite number of units of a given size Past Greek conceptions dictated that they necessarily must be for whole numbers represent discrete objects and a commensurable ratio represents a relation between two collections of discrete objects 8 but Zeno found that in fact quantities in general are not discrete collections of units this is why ratios of incommensurable quantities appear Q uantities are in other words continuous 8 What this means is that contrary to the popular conception of the time there cannot be an indivisible smallest unit of measure for any quantity In fact these divisions of quantity must necessarily be infinite For example consider a line segment this segment can be split in half that half split in half the half of the half in half and so on This process can continue infinitely for there is always another half to be split The more times the segment is halved the closer the unit of measure comes to zero but it never reaches exactly zero This is just what Zeno sought to prove He sought to prove this by formulating four paradoxes which demonstrated the contradictions inherent in the mathematical thought of the time While Zeno s paradoxes accurately demonstrated the deficiencies of contemporary mathematical conceptions they were not regarded as proof of the alternative In the minds of the Greeks disproving the validity of one view did not necessarily prove the validity of another and therefore further investigation had to occur The next step was taken by Eudoxus of Cnidus who formalized a new theory of proportion that took into account commensurable as well as incommensurable quantities Central to his idea was the distinction between magnitude and number A magnitude was not a number but stood for entities such as line segments angles areas volumes and time which could vary as we would say continuously Magnitudes were opposed to numbers which jumped from one value to another as from 4 to 5 9 Numbers are composed of some smallest indivisible unit whereas magnitudes are infinitely reducible Because no quantitative values were assigned to magnitudes Eudoxus was then able to account for both commensurable and incommensurable ratios by defining a ratio in terms of its magnitude and proportion as an equality between two ratios By taking quantitative values numbers out of the equation he avoided the trap of having to express an irrational number as a number Eudoxus theory enabled the Greek mathematicians to make tremendous progress in geometry by supplying the necessary logical foundation for incommensurable ratios 10 This incommensurability is dealt with in Euclid s Elements Book X Proposition 9 It was not until Eudoxus developed a theory of proportion that took into account irrational as well as rational ratios that a strong mathematical foundation of irrational numbers was created 11 As a result of the distinction between number and magnitude geometry became the only method that could take into account incommensurable ratios Because previous numerical foundations were still incompatible with the concept of incommensurability Greek focus shifted away from numerical conceptions such as algebra and focused almost exclusively on geometry In fact in many cases algebraic conceptions were reformulated into geometric terms This may account for why we still conceive of x2 and x3 as x squared and x cubed instead of x to the second power and x to the third power Also crucial to Zeno s work with incommensurable magnitudes was the fundamental focus on deductive reasoning that resulted from the foundational shattering of earlier Greek mathematics The realization that some basic conception within the existing theory was at odds with reality necessitated a complete and thorough investigation of the axioms and assumptions that underlie that theory Out of this necessity Eudoxus developed his method of exhaustion a kind of reductio ad absurdum that established the deductive organization on the basis of explicit axioms as well as reinforced the earlier decision to rely on deductive reasoning for proof 12 This method of exhaustion is the first step in the creation of calculus Theodorus of Cyrene proved the irrationality of the surds of whole numbers up to 17 but stopped there probably because the algebra he used could not be applied to the square root of 17 13 India edit Geometrical and mathematical problems involving irrational numbers such as square roots were addressed very early during the Vedic period in India There are references to such calculations in the Samhitas Brahmanas and the Shulba Sutras 800 BC or earlier 14 It is suggested that the concept of irrationality was implicitly accepted by Indian mathematicians since the 7th century BC when Manava c 750 690 BC believed that the square roots of numbers such as 2 and 61 could not be exactly determined 15 Historian Carl Benjamin Boyer however writes that such claims are not well substantiated and unlikely to be true 16 Later in their treatises Indian mathematicians wrote on the arithmetic of surds including addition subtraction multiplication rationalization as well as separation and extraction of square roots 17 Mathematicians like Brahmagupta in 628 AD and Bhaskara I in 629 AD made contributions in this area as did other mathematicians who followed In the 12th century Bhaskara II evaluated some of these formulas and critiqued them identifying their limitations During the 14th to 16th centuries Madhava of Sangamagrama and the Kerala school of astronomy and mathematics discovered the infinite series for several irrational numbers such as p and certain irrational values of trigonometric functions Jyeṣṭhadeva provided proofs for these infinite series in the Yuktibhaṣa 18 Islamic World edit In the Middle Ages the development of algebra by Muslim mathematicians allowed irrational numbers to be treated as algebraic objects 19 Middle Eastern mathematicians also merged the concepts of number and magnitude into a more general idea of real numbers criticized Euclid s idea of ratios developed the theory of composite ratios and extended the concept of number to ratios of continuous magnitude 20 In his commentary on Book 10 of the Elements the Persian mathematician Al Mahani d 874 884 examined and classified quadratic irrationals and cubic irrationals He provided definitions for rational and irrational magnitudes which he treated as irrational numbers He dealt with them freely but explains them in geometric terms as follows 20 It will be a rational magnitude when we for instance say 10 12 3 6 etc because its value is pronounced and expressed quantitatively What is not rational is irrational and it is impossible to pronounce and represent its value quantitatively For example the roots of numbers such as 10 15 20 which are not squares the sides of numbers which are not cubes etc In contrast to Euclid s concept of magnitudes as lines Al Mahani considered integers and fractions as rational magnitudes and square roots and cube roots as irrational magnitudes He also introduced an arithmetical approach to the concept of irrationality as he attributes the following to irrational magnitudes 20 their sums or differences or results of their addition to a rational magnitude or results of subtracting a magnitude of this kind from an irrational one or of a rational magnitude from it The Egyptian mathematician Abu Kamil Shuja ibn Aslam c 850 930 was the first to accept irrational numbers as solutions to quadratic equations or as coefficients in an equation in the form of square roots and fourth roots 21 In the 10th century the Iraqi mathematician Al Hashimi provided general proofs rather than geometric demonstrations for irrational numbers as he considered multiplication division and other arithmetical functions 20 Many of these concepts were eventually accepted by European mathematicians sometime after the Latin translations of the 12th century Al Hassar a Moroccan mathematician from Fez specializing in Islamic inheritance jurisprudence during the 12th century first mentions the use of a fractional bar where numerators and denominators are separated by a horizontal bar In his discussion he writes for example if you are told to write three fifths and a third of a fifth write thus 3 1 5 3 displaystyle frac 3 quad 1 5 quad 3 nbsp 22 This same fractional notation appears soon after in the work of Leonardo Fibonacci in the 13th century 23 Modern period edit The 17th century saw imaginary numbers become a powerful tool in the hands of Abraham de Moivre and especially of Leonhard Euler The completion of the theory of complex numbers in the 19th century entailed the differentiation of irrationals into algebraic and transcendental numbers the proof of the existence of transcendental numbers and the resurgence of the scientific study of the theory of irrationals largely ignored since Euclid The year 1872 saw the publication of the theories of Karl Weierstrass by his pupil Ernst Kossak Eduard Heine Crelle s Journal 74 Georg Cantor Annalen 5 and Richard Dedekind Meray had taken in 1869 the same point of departure as Heine but the theory is generally referred to the year 1872 Weierstrass s method has been completely set forth by Salvatore Pincherle in 1880 24 and Dedekind s has received additional prominence through the author s later work 1888 and the endorsement by Paul Tannery 1894 Weierstrass Cantor and Heine base their theories on infinite series while Dedekind founds his on the idea of a cut Schnitt in the system of all rational numbers separating them into two groups having certain characteristic properties The subject has received later contributions at the hands of Weierstrass Leopold Kronecker Crelle 101 and Charles Meray Continued fractions closely related to irrational numbers and due to Cataldi 1613 received attention at the hands of Euler and at the opening of the 19th century were brought into prominence through the writings of Joseph Louis Lagrange Dirichlet also added to the general theory as have numerous contributors to the applications of the subject Johann Heinrich Lambert proved 1761 that p cannot be rational and that en is irrational if n is rational unless n 0 25 While Lambert s proof is often called incomplete modern assessments support it as satisfactory and in fact for its time it is unusually rigorous Adrien Marie Legendre 1794 after introducing the Bessel Clifford function provided a proof to show that p2 is irrational whence it follows immediately that p is irrational also The existence of transcendental numbers was first established by Liouville 1844 1851 Later Georg Cantor 1873 proved their existence by a different method which showed that every interval in the reals contains transcendental numbers Charles Hermite 1873 first proved e transcendental and Ferdinand von Lindemann 1882 starting from Hermite s conclusions showed the same for p Lindemann s proof was much simplified by Weierstrass 1885 still further by David Hilbert 1893 and was finally made elementary by Adolf Hurwitz citation needed and Paul Gordan 26 Examples editThis article needs additional citations for verification Please help improve this article by adding citations to reliable sources Unsourced material may be challenged and removed Find sources Irrational number news newspapers books scholar JSTOR May 2023 Learn how and when to remove this message Square roots edit The square root of 2 was likely the first number proved irrational 27 The golden ratio is another famous quadratic irrational number The square roots of all natural numbers that are not perfect squares are irrational and a proof may be found in quadratic irrationals General roots edit The proof above clarification needed for the square root of two can be generalized using the fundamental theorem of arithmetic This asserts that every integer has a unique factorization into primes Using it we can show that if a rational number is not an integer then no integral power of it can be an integer as in lowest terms there must be a prime in the denominator that does not divide into the numerator whatever power each is raised to Therefore if an integer is not an exact k th power of another integer then that first integer s k th root is irrational Logarithms edit Perhaps the numbers most easy to prove irrational are certain logarithms Here is a proof by contradiction that log2 3 is irrational log2 3 1 58 gt 0 Assume log2 3 is rational For some positive integers m and n we have log 2 3 m n displaystyle log 2 3 frac m n nbsp It follows that 2 m n 3 displaystyle 2 m n 3 nbsp 2 m n n 3 n displaystyle 2 m n n 3 n nbsp 2 m 3 n displaystyle 2 m 3 n nbsp The number 2 raised to any positive integer power must be even because it is divisible by 2 and the number 3 raised to any positive integer power must be odd since none of its prime factors will be 2 Clearly an integer cannot be both odd and even at the same time we have a contradiction The only assumption we made was that log2 3 is rational and so expressible as a quotient of integers m n with n 0 The contradiction means that this assumption must be false i e log2 3 is irrational and can never be expressed as a quotient of integers m n with n 0 Cases such as log10 2 can be treated similarly Types editAn irrational number may be algebraic that is a real root of a polynomial with integer coefficients Those that are not algebraic are transcendental Algebraic edit The real algebraic numbers are the real solutions of polynomial equations p x a n x n a n 1 x n 1 a 1 x a 0 0 displaystyle p x a n x n a n 1 x n 1 cdots a 1 x a 0 0 nbsp where the coefficients a i displaystyle a i nbsp are integers and a n 0 displaystyle a n neq 0 nbsp An example of an irrational algebraic number is x0 21 2 1 1 3 It is clearly algebraic since it is the root of an integer polynomial x3 1 2 2 which is equivalent to x6 2x3 1 0 This polynomial has no rational roots since the rational root theorem shows that the only possibilities are 1 but x0is greater than 1 So x0 is an irrational algebraic number There are countably many algebraic numbers since there are countably many integer polynomials Transcendental edit So almost all irrational numbers are transcendental Examples are e r and p r which are transcendental for all nonzero rational r Because the algebraic numbers form a subfield of the real numbers many irrational real numbers can be constructed by combining transcendental and algebraic numbers For example 3p 2 p 2 and e 3 are irrational and even transcendental Decimal expansions editThe decimal expansion of an irrational number never repeats meaning the decimal expansion does not repeat the same number or sequence of numbers or terminates this means there is not a finite number of nonzero digits unlike any rational number The same is true for binary octal or hexadecimal expansions and in general for expansions in every positional notation with natural bases To show this suppose we divide integers n by m where m is nonzero When long division is applied to the division of n by m there can never be a remainder greater than or equal to m If 0 appears as a remainder the decimal expansion terminates If 0 never occurs then the algorithm can run at most m 1 steps without using any remainder more than once After that a remainder must recur and then the decimal expansion repeats Conversely suppose we are faced with a repeating decimal we can prove that it is a fraction of two integers For example consider A 0 7 162 162 162 displaystyle A 0 7 162 162 162 ldots nbsp Here the repetend is 162 and the length of the repetend is 3 First we multiply by an appropriate power of 10 to move the decimal point to the right so that it is just in front of a repetend In this example we would multiply by 10 to obtain 10 A 7 162 162 162 displaystyle 10A 7 162 162 162 ldots nbsp Now we multiply this equation by 10r where r is the length of the repetend This has the effect of moving the decimal point to be in front of the next repetend In our example multiply by 103 10 000 A 7 162 162 162 displaystyle 10 000A 7 162 162 162 ldots nbsp The result of the two multiplications gives two different expressions with exactly the same decimal portion that is the tail end of 10 000A matches the tail end of 10A exactly Here both 10 000A and 10A have 162162 162 after the decimal point Therefore when we subtract the 10A equation from the 10 000A equation the tail end of 10A cancels out the tail end of 10 000A leaving us with 9990 A 7155 displaystyle 9990A 7155 nbsp Then A 7155 9990 53 74 displaystyle A frac 7155 9990 frac 53 74 nbsp is a ratio of integers and therefore a rational number Irrational powers editDov Jarden gave a simple non constructive proof that there exist two irrational numbers a and b such that ab is rational 28 29 Consider 2 2 if this is rational then take a b 2 Otherwise take a to be the irrational number 2 2 and b 2 Then ab 2 2 2 2 2 2 2 2 2 which is rational Although the above argument does not decide between the two cases the Gelfond Schneider theorem shows that 2 2 is transcendental hence irrational This theorem states that if a and b are both algebraic numbers and a is not equal to 0 or 1 and b is not a rational number then any value of ab is a transcendental number there can be more than one value if complex number exponentiation is used An example that provides a simple constructive proof is 30 2 log 2 3 3 displaystyle left sqrt 2 right log sqrt 2 3 3 nbsp The base of the left side is irrational and the right side is rational so one must prove that the exponent on the left side log 2 3 displaystyle log sqrt 2 3 nbsp is irrational This is so because by the formula relating logarithms with different bases log 2 3 log 2 3 log 2 2 log 2 3 1 2 2 log 2 3 displaystyle log sqrt 2 3 frac log 2 3 log 2 sqrt 2 frac log 2 3 1 2 2 log 2 3 nbsp which we can assume for the sake of establishing a contradiction equals a ratio m n of positive integers Then log 2 3 m 2 n displaystyle log 2 3 m 2n nbsp hence 2 log 2 3 2 m 2 n displaystyle 2 log 2 3 2 m 2n nbsp hence 3 2 m 2 n displaystyle 3 2 m 2n nbsp hence 3 2 n 2 m displaystyle 3 2n 2 m nbsp which is a contradictory pair of prime factorizations and hence violates the fundamental theorem of arithmetic unique prime factorization A stronger result is the following 31 Every rational number in the interval 1 e 1 e displaystyle 1 e 1 e infty nbsp can be written either as aa for some irrational number a or as nn for some natural number n Similarly 31 every positive rational number can be written either as a a a displaystyle a a a nbsp for some irrational number a or as n n n displaystyle n n n nbsp for some natural number n Open questions editIt is not known if p e displaystyle pi e nbsp or p e displaystyle pi e nbsp is irrational In fact there is no pair of non zero integers m n displaystyle m n nbsp for which it is known whether m p n e displaystyle m pi ne nbsp is irrational Moreover it is not known if the set p e displaystyle pi e nbsp is algebraically independent over Q displaystyle mathbb Q nbsp It is not known if p e p e p e p 2 ln p displaystyle pi e pi e pi e pi sqrt 2 ln pi nbsp Catalan s constant or the Euler Mascheroni constant g displaystyle gamma nbsp are irrational 32 It is not known if either of the tetrations n p displaystyle n pi nbsp or n e displaystyle n e nbsp is rational for some integer n gt 1 displaystyle n gt 1 nbsp citation needed In constructive mathematics editIn constructive mathematics excluded middle is not valid so it is not true that every real number is rational or irrational Thus the notion of an irrational number bifurcates into multiple distinct notions One could take the traditional definition of an irrational number as a real number that is not rational 33 However there is a second definition of an irrational number used in constructive mathematics that a real number r displaystyle r nbsp is an irrational number if it is apart from every rational number or equivalently if the distance r q displaystyle vert r q vert nbsp between r displaystyle r nbsp and every rational number q displaystyle q nbsp is positive This definition is stronger than the traditional definition of an irrational number This second definition is used in Errett Bishop s proof that the square root of 2 is irrational 34 Set of all irrationals editSince the reals form an uncountable set of which the rationals are a countable subset the complementary set of irrationals is uncountable Under the usual Euclidean distance function d x y x y displaystyle d x y vert x y vert nbsp the real numbers are a metric space and hence also a topological space Restricting the Euclidean distance function gives the irrationals the structure of a metric space Since the subspace of irrationals is not closed the induced metric is not complete Being a G delta set i e a countable intersection of open subsets in a complete metric space the space of irrationals is completely metrizable that is there is a metric on the irrationals inducing the same topology as the restriction of the Euclidean metric but with respect to which the irrationals are complete One can see this without knowing the aforementioned fact about G delta sets the continued fraction expansion of an irrational number defines a homeomorphism from the space of irrationals to the space of all sequences of positive integers which is easily seen to be completely metrizable Furthermore the set of all irrationals is a disconnected metrizable space In fact the irrationals equipped with the subspace topology have a basis of clopen groups so the space is zero dimensional See also editBrjuno number Computable number Diophantine approximation Proof that e is irrational Proof that p is irrational Square root of 3 Square root of 5 Trigonometric number Number systems Complex C displaystyle mathbb C nbsp Real R displaystyle mathbb R nbsp Rational Q displaystyle mathbb Q nbsp Integer Z displaystyle mathbb Z nbsp Natural N displaystyle mathbb N nbsp Zero 0 One 1 Prime numbers Composite numbers Negative integers Fraction Finite decimal Dyadic finite binary Repeating decimal Irrational Algebraic irrational Transcendental ImaginaryReferences edit The 15 Most Famous Transcendental Numbers by Clifford A Pickover URL retrieved 24 October 2007 Jackson Terence 2011 07 01 95 42 Irrational square roots of natural numbers a geometrical approach The Mathematical Gazette 95 533 327 330 doi 10 1017 S0025557200003193 ISSN 0025 5572 S2CID 123995083 Cantor Georg 1955 1915 Philip Jourdain ed Contributions to the Founding of the Theory of Transfinite Numbers New York Dover ISBN 978 0 486 60045 1 Kurt Von Fritz 1945 The Discovery of Incommensurability by Hippasus of Metapontum Annals of Mathematics 46 2 242 264 doi 10 2307 1969021 JSTOR 1969021 S2CID 126296119 James R Choike 1980 The Pentagram and the Discovery of an Irrational Number The Two Year College Mathematics Journal 11 5 312 316 doi 10 2307 3026893 JSTOR 3026893 S2CID 115390951 Kline M 1990 Mathematical Thought from Ancient to Modern Times Vol 1 New York Oxford University Press original work published 1972 p 33 Kline 1990 p 32 a b Kline 1990 p 34 Kline 1990 p 48 Kline 1990 p 49 Charles H Edwards 1982 The historical development of the calculus Springer Kline 1990 p 50 Robert L McCabe 1976 Theodorus Irrationality Proofs Mathematics Magazine 49 4 201 203 doi 10 1080 0025570X 1976 11976579 JSTOR 2690123 S2CID 124565880 Bag Amulya Kumar 1990 Ritual Geometry in India and its Parallelism in other Culture Areas Indian Journal of History of Science 25 T K Puttaswamy The Accomplishments of Ancient Indian Mathematicians pp 411 2 in Selin Helaine D Ambrosio Ubiratan eds 2000 Mathematics Across Cultures The History of Non western Mathematics Springer ISBN 1 4020 0260 2 Boyer 1991 China and India A History of Mathematics 2nd ed Wiley p 208 ISBN 0471093742 OCLC 414892 It has been claimed also that the first recognition of incommensurables appears in India during the Sulbasutra period but such claims are not well substantiated The case for early Hindu awareness of incommensurable magnitudes is rendered most unlikely by the lack of evidence that Indian mathematicians of that period had come to grips with fundamental concepts Datta Bibhutibhusan Singh Awadhesh Narayan 1993 Surds in Hindu mathematics PDF Indian Journal of History of Science 28 3 253 264 Archived from the original PDF on 2018 10 03 Retrieved 18 September 2018 Katz V J 1995 Ideas of Calculus in Islam and India Mathematics Magazine 63 3 163 174 doi 10 2307 2691411 JSTOR 2691411 O Connor John J Robertson Edmund F 1999 Arabic mathematics forgotten brilliance MacTutor History of Mathematics Archive University of St Andrews a b c d Matvievskaya Galina 1987 The theory of quadratic irrationals in medieval Oriental mathematics Annals of the New York Academy of Sciences 500 1 253 277 Bibcode 1987NYASA 500 253M doi 10 1111 j 1749 6632 1987 tb37206 x S2CID 121416910 See in particular pp 254 amp 259 260 Jacques Sesiano Islamic mathematics p 148 in Selin Helaine D Ambrosio Ubiratan 2000 Mathematics Across Cultures The History of Non western Mathematics Springer ISBN 1 4020 0260 2 Cajori Florian 1928 A History of Mathematical Notations Vol 1 La Salle Illinois The Open Court Publishing Company pg 269 Cajori 1928 pg 89 Salvatore Pincherle 1880 Saggio di una introduzione alla teoria delle funzioni analitiche secondo i principii del prof C Weierstrass Giornale di Matematiche 178 254 317 320 Lambert J H 1761 Memoire sur quelques proprietes remarquables des quantites transcendentes circulaires et logarithmiques PDF Memoires de l Academie royale des sciences de Berlin in French 265 322 Archived PDF from the original on 2016 04 28 Gordan Paul 1893 Transcendenz von e und p Mathematische Annalen 43 2 3 Teubner 222 224 doi 10 1007 bf01443647 S2CID 123203471 Fowler David H 2001 The story of the discovery of incommensurability revisited Neusis 10 45 61 MR 1891736 Jarden Dov 1953 Curiosa No 339 A simple proof that a power of an irrational number to an irrational exponent may be rational Scripta Mathematica 19 229 copy George Alexander Velleman Daniel J 2002 Philosophies of mathematics PDF Blackwell pp 3 4 ISBN 0 631 19544 0 Lord Nick Maths bite irrational powers of irrational numbers can be rational Mathematical Gazette 92 November 2008 p 534 a b Marshall Ash J and Tan Yiren A rational number of the form aa with a irrational Mathematical Gazette 96 March 2012 pp 106 109 Albert John Some unsolved problems in number theory PDF Department of Mathematics University of Oklahoma Senior Mathematics Seminar Spring 2008 course Mark Bridger 2007 Real Analysis A Constructive Approach through Interval Arithmetic John Wiley amp Sons ISBN 978 1 470 45144 8 Errett Bishop Douglas Bridges 1985 Constructive Analysis Springer ISBN 0 387 15066 8 Further reading editAdrien Marie Legendre Elements de Geometrie Note IV 1802 Paris Rolf Wallisser On Lambert s proof of the irrationality of p in Algebraic Number Theory and Diophantine Analysis Franz Halter Koch and Robert F Tichy 2000 Walter de GruyterExternal links edit nbsp Wikimedia Commons has media related to Irrational numbers Zeno s Paradoxes and Incommensurability Archived 2016 05 13 at the Wayback Machine n d Retrieved April 1 2008 Retrieved from https en wikipedia org w index php title Irrational number amp oldid 1217181007, 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.