Jump to content
 







Main menu
   


Navigation  



Main page
Contents
Current events
Random article
About Wikipedia
Contact us
Donate
 




Contribute  



Help
Learn to edit
Community portal
Recent changes
Upload file
 








Search  

































Create account

Log in
 









Create account
 Log in
 




Pages for logged out editors learn more  



Contributions
Talk
 



















Contents

   



(Top)
 


1 Lambert's proof  





2 Hermite's proof  





3 Cartwright's proof  





4 Niven's proof  





5 Bourbaki's proof  





6 Laczkovich's proof  





7 See also  





8 References  














Proof that π is irrational






العربية
Español
فارسی
Français

Italiano
עברית
Magyar

ି
Português
Русский
Українська
Tiếng Vit

 

Edit links
 









Article
Talk
 

















Read
Edit
View history
 








Tools
   


Actions  



Read
Edit
View history
 




General  



What links here
Related changes
Upload file
Special pages
Permanent link
Page information
Cite this page
Get shortened URL
Download QR code
Wikidata item
 




Print/export  



Download as PDF
Printable version
 
















Appearance
   

 






From Wikipedia, the free encyclopedia
 


In the 1760s, Johann Heinrich Lambert was the first to prove that the number πisirrational, meaning it cannot be expressed as a fraction , where and are both integers. In the 19th century, Charles Hermite found a proof that requires no prerequisite knowledge beyond basic calculus. Three simplifications of Hermite's proof are due to Mary Cartwright, Ivan Niven, and Nicolas Bourbaki. Another proof, which is a simplification of Lambert's proof, is due to Miklós Laczkovich. Many of these are proofs by contradiction.

In 1882, Ferdinand von Lindemann proved that is not just irrational, but transcendental as well.[1]

Lambert's proof[edit]

Scan of formula on page 288 of Lambert's "Mémoires sur quelques propriétés remarquables des quantités transcendantes, circulaires et logarithmiques", Mémoires de l'Académie royale des sciences de Berlin (1768), 265–322

In 1761, Johann Heinrich Lambert proved that is irrational by first showing that this continued fraction expansion holds:

Then Lambert proved that if is non-zero and rational, then this expression must be irrational. Since , it follows that is irrational, and thus is also irrational.[2] A simplification of Lambert's proof is given below.

Hermite's proof[edit]

Written in 1873, this proof uses the characterization of as the smallest positive number whose half is a zero of the cosine function and it actually proves that is irrational.[3][4] As in many proofs of irrationality, it is a proof by contradiction.

Consider the sequences of real functions and for defined by:

Using induction we can prove that

and therefore we have:

So

which is equivalent to

Using the definition of the sequence and employing induction we can show that

where and are polynomial functions with integer coefficients and the degree of is smaller than or equal to In particular,

Hermite also gave a closed expression for the function namely

He did not justify this assertion, but it can be proved easily. First of all, this assertion is equivalent to

Proceeding by induction, take

and, for the inductive step, consider any natural number If

then, using integration by parts and Leibniz's rule, one gets

If with and in, then, since the coefficients of are integers and its degree is smaller than or equal to is some integer In other words,

But this number is clearly greater than On the other hand, the limit of this quantity as goes to infinity is zero, and so, if is large enough, Thereby, a contradiction is reached.

Hermite did not present his proof as an end in itself but as an afterthought within his search for a proof of the transcendence of He discussed the recurrence relations to motivate and to obtain a convenient integral representation. Once this integral representation is obtained, there are various ways to present a succinct and self-contained proof starting from the integral (as in Cartwright's, Bourbaki's or Niven's presentations), which Hermite could easily see (as he did in his proof of the transcendence of [5]).

Moreover, Hermite's proof is closer to Lambert's proof than it seems. In fact, is the "residue" (or "remainder") of Lambert's continued fraction for [6]

Cartwright's proof[edit]

Harold Jeffreys wrote that this proof was set as an example in an exam at Cambridge University in 1945 by Mary Cartwright, but that she had not traced its origin.[7] It still remains on the 4th problem sheet today for the Analysis IA course at Cambridge University.[8]

Consider the integrals

where is a non-negative integer.

Two integrations by parts give the recurrence relation

If

then this becomes

Furthermore, and Hence for all

where and are polynomials of degree and with integer coefficients (depending on ).

Take and suppose if possible that where and are natural numbers (i.e., assume that is rational). Then

The right side is an integer. But since the interval has length and the function being integrated takes only values between and On the other hand,

Hence, for sufficiently large

that is, we could find an integer between and That is the contradiction that follows from the assumption that is rational.

This proof is similar to Hermite's proof. Indeed,

However, it is clearly simpler. This is achieved by omitting the inductive definition of the functions and taking as a starting point their expression as an integral.

Niven's proof[edit]

This proof uses the characterization of as the smallest positive zero of the sine function.[9]

Suppose that is rational, i.e. for some integers and which may be taken without loss of generality to both be positive. Given any positive integer we define the polynomial function:

and, for each let

Claim 1: is an integer.

Proof: Expanding as a sum of monomials, the coefficient of is a number of the form where is an integer, which is if Therefore, is when and it is equal to if; in each case, is an integer and therefore is an integer.

On the other hand, and so for each non-negative integer In particular, Therefore, is also an integer and so is an integer (in fact, it is easy to see that ). Since and are integers, so is their sum.

Claim 2:

Proof: Since is the zero polynomial, we have

The derivatives of the sine and cosine function are given by sin' = cos and cos' = −sin. Hence the product rule implies

By the fundamental theorem of calculus

Since and (here we use the above-mentioned characterization of as a zero of the sine function), Claim 2 follows.

Conclusion: Since and for (because is the smallest positive zero of the sine function), Claims 1 and 2 show that is a positive integer. Since and for we have, by the original definition of

which is smaller than for large hence for these by Claim 2. This is impossible for the positive integer This shows that the original assumption that is rational leads to a contradiction, which concludes the proof.

The above proof is a polished version, which is kept as simple as possible concerning the prerequisites, of an analysis of the formula

which is obtained by integrations by parts. Claim 2 essentially establishes this formula, where the use of hides the iterated integration by parts. The last integral vanishes because is the zero polynomial. Claim 1 shows that the remaining sum is an integer.

Niven's proof is closer to Cartwright's (and therefore Hermite's) proof than it appears at first sight.[6] In fact,

Therefore, the substitution turns this integral into

In particular,

Another connection between the proofs lies in the fact that Hermite already mentions[3] that if is a polynomial function and

then

from which it follows that

Bourbaki's proof[edit]

Bourbaki's proof is outlined as an exercise in his calculus treatise.[10] For each natural number b and each non-negative integer define

Since is the integral of a function defined on that takes the value at and and which is greater than otherwise, Besides, for each natural number if is large enough, because

and therefore

On the other hand, repeated integration by parts allows us to deduce that, if and are natural numbers such that and is the polynomial function from into defined by

then:

This last integral is since is the null function (because is a polynomial function of degree ). Since each function (with ) takes integer values at and and since the same thing happens with the sine and the cosine functions, this proves that is an integer. Since it is also greater than it must be a natural number. But it was also proved that if is large enough, thereby reaching a contradiction.

This proof is quite close to Niven's proof, the main difference between them being the way of proving that the numbers are integers.

Laczkovich's proof[edit]

Miklós Laczkovich's proof is a simplification of Lambert's original proof.[11] He considers the functions

These functions are clearly defined for any real number Besides

Claim 1: The following recurrence relation holds for any real number :

Proof: This can be proved by comparing the coefficients of the powers of

Claim 2: For each real number

Proof: In fact, the sequence is bounded (since it converges to ) and if is an upper bound and if then

Claim 3:If is rational, and then

Proof: Otherwise, there would be a number and integers and such that and To see why, take and if; otherwise, choose integers and such that and define In each case, cannot be because otherwise it would follow from claim 1 that each () would be which would contradict claim 2. Now, take a natural number such that all three numbers and are integers and consider the sequence

Then

On the other hand, it follows from claim 1 that

which is a linear combination of and with integer coefficients. Therefore, each is an integer multiple of Besides, it follows from claim 2 that each is greater than (and therefore that )if is large enough and that the sequence of all converges to But a sequence of numbers greater than or equal to cannot converge to

Since it follows from claim 3 that is irrational and therefore that is irrational.

On the other hand, since

another consequence of Claim 3 is that, if then is irrational.

Laczkovich's proof is really about the hypergeometric function. In fact, and Gauss found a continued fraction expansion of the hypergeometric function using its functional equation.[12] This allowed Laczkovich to find a new and simpler proof of the fact that the tangent function has the continued fraction expansion that Lambert had discovered.

Laczkovich's result can also be expressed in Bessel functions of the first kind . In fact, (where is the gamma function). So Laczkovich's result is equivalent to: If is rational, and then

See also[edit]

References[edit]

  1. ^ Lindemann, Ferdinand von (2004) [1882], "Ueber die Zahl π", in Berggren, Lennart; Borwein, Jonathan M.; Borwein, Peter B. (eds.), Pi, a source book (3rd ed.), New York: Springer-Verlag, pp. 194–225, ISBN 0-387-20571-3.
  • ^ Lambert, Johann Heinrich (2004) [1768], "Mémoire sur quelques propriétés remarquables des quantités transcendantes circulaires et logarithmiques", in Berggren, Lennart; Borwein, Jonathan M.; Borwein, Peter B. (eds.), Pi, a source book (3rd ed.), New York: Springer-Verlag, pp. 129–140, ISBN 0-387-20571-3.
  • ^ a b Hermite, Charles (1873). "Extrait d'une lettre de Monsieur Ch. Hermite à Monsieur Paul Gordan". Journal für die reine und angewandte Mathematik (in French). 76: 303–311.
  • ^ Hermite, Charles (1873). "Extrait d'une lettre de Mr. Ch. Hermite à Mr. Carl Borchardt". Journal für die reine und angewandte Mathematik (in French). 76: 342–344.
  • ^ Hermite, Charles (1912) [1873]. "Sur la fonction exponentielle". In Picard, Émile (ed.). Œuvres de Charles Hermite (in French). Vol. III. Gauthier-Villars. pp. 150–181.
  • ^ a b Zhou, Li (2011). "Irrationality proofs à la Hermite". The Mathematical Gazette. 95 (534): 407–413. arXiv:0911.1929. doi:10.1017/S0025557200003491. S2CID 115175505.
  • ^ Jeffreys, Harold (1973), Scientific Inference (3rd ed.), Cambridge University Press, p. 268, ISBN 0-521-08446-6
  • ^ "Department of Pure Mathematics and Mathematical Statistics". www.dpmms.cam.ac.uk. Retrieved 2022-04-19.
  • ^ Niven, Ivan (1947), "A simple proof that π is irrational" (PDF), Bulletin of the American Mathematical Society, vol. 53, no. 6, p. 509, doi:10.1090/s0002-9904-1947-08821-2
  • ^ Bourbaki, Nicolas (1949), Fonctions d'une variable réelle, chap. I–II–III, Actualités Scientifiques et Industrielles (in French), vol. 1074, Hermann, pp. 137–138
  • ^ Laczkovich, Miklós (1997), "On Lambert's proof of the irrationality of π", American Mathematical Monthly, vol. 104, no. 5, pp. 439–443, doi:10.2307/2974737, JSTOR 2974737
  • ^ Gauss, Carl Friedrich (1811–1813), "Disquisitiones generales circa seriem infinitam", Commentationes Societatis Regiae Scientiarum Gottingensis Recentiores (in Latin), 2

  • Retrieved from "https://en.wikipedia.org/w/index.php?title=Proof_that_π_is_irrational&oldid=1218938485"

    Categories: 
    Pi
    Article proofs
    Irrational numbers
    Hidden categories: 
    CS1 French-language sources (fr)
    CS1 Latin-language sources (la)
    Articles with short description
    Short description matches Wikidata
     



    This page was last edited on 14 April 2024, at 19:49 (UTC).

    Text is available under the Creative Commons Attribution-ShareAlike License 4.0; additional terms may apply. By using this site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc., a non-profit organization.



    Privacy policy

    About Wikipedia

    Disclaimers

    Contact Wikipedia

    Code of Conduct

    Developers

    Statistics

    Cookie statement

    Mobile view



    Wikimedia Foundation
    Powered by MediaWiki