fbpx
Wikipedia

Riemann series theorem

In mathematics, the Riemann series theorem, also called the Riemann rearrangement theorem, named after 19th-century German mathematician Bernhard Riemann, says that if an infinite series of real numbers is conditionally convergent, then its terms can be arranged in a permutation so that the new series converges to an arbitrary real number, or diverges. This implies that a series of real numbers is absolutely convergent if and only if it is unconditionally convergent.

As an example, the series 1 − 1 + 1/2 − 1/2 + 1/3 − 1/3 + ⋯ converges to 0 (for a sufficiently large number of terms, the partial sum gets arbitrarily near to 0); but replacing all terms with their absolute values gives 1 + 1 + 1/2 + 1/2 + 1/3 + 1/3 + ⋯, which sums to infinity. Thus the original series is conditionally convergent, and can be rearranged (by taking the first two positive terms followed by the first negative term, followed by the next two positive terms and then the next negative term, etc.) to give a series that converges to a different sum: 1 + 1/2 − 1 + 1/3 + 1/4 − 1/2 + ⋯ = ln 2. More generally, using this procedure with p positives followed by q negatives gives the sum ln(p/q). Other rearrangements give other finite sums or do not converge to any sum.

History edit

It is a basic result that the sum of finitely many numbers does not depend on the order in which they are added. For example, 2 + 6 + 7 = 7 + 2 + 6. The observation that the sum of an infinite sequence of numbers can depend on the ordering of the summands is commonly attributed to Augustin-Louis Cauchy in 1833.[1] He analyzed the alternating harmonic series, showing that certain rearrangements of its summands result in different limits. Around the same time, Peter Gustav Lejeune Dirichlet highlighted that such phenomena are ruled out in the context of absolute convergence, and gave further examples of Cauchy's phenomenon for some other series which fail to be absolutely convergent.[2]

In the course of his analysis of Fourier series and the theory of Riemann integration, Bernhard Riemann gave a full characterization of the rearrangement phenomena.[3] He proved that in the case of a convergent series which does not converge absolutely (known as conditional convergence), rearrangements can be found so that the new series converges to any arbitrarily prescribed real number.[4] Riemann's theorem is now considered as a basic part of the field of mathematical analysis.[5]

For any series, one may consider the set of all possible sums, corresponding to all possible rearrangements of the summands. Riemann’s theorem can be formulated as saying that, for a series of real numbers, this set is either empty, a single point (in the case of absolute convergence), or the entire real number line (in the case of conditional convergence). In this formulation, Riemann’s theorem was extended by Paul Lévy and Ernst Steinitz to series whose summands are complex numbers or, even more generally, elements of a finite-dimensional real vector space. They proved that the set of possible sums forms a real affine subspace. Extensions of the Lévy–Steinitz theorem to series in infinite-dimensional spaces have been considered by a number of authors.[6]

Definitions edit

A series   converges if there exists a value   such that the sequence of the partial sums

 

converges to  . That is, for any ε > 0, there exists an integer N such that if n ≥ N, then

 

A series converges conditionally if the series   converges but the series   diverges.

A permutation is simply a bijection from the set of positive integers to itself. This means that if   is a permutation, then for any positive integer   there exists exactly one positive integer   such that   In particular, if  , then  .

Statement of the theorem edit

Suppose that   is a sequence of real numbers, and that   is conditionally convergent. Let   be a real number. Then there exists a permutation   such that

 

There also exists a permutation   such that

 

The sum can also be rearranged to diverge to   or to fail to approach any limit, finite or infinite.

Alternating harmonic series edit

Changing the sum edit

The alternating harmonic series is a classic example of a conditionally convergent series:

 
is convergent, whereas
 
is the ordinary harmonic series, which diverges. Although in standard presentation the alternating harmonic series converges to ln(2), its terms can be arranged to converge to any number, or even to diverge.


One instance of this is as follows. Begin with the series written in the usual order,

 

and rearrange and regroup the terms as:

 

where the pattern is: the first two terms are 1 and −1/2, whose sum is 1/2. The next term is −1/4. The next two terms are 1/3 and −1/6, whose sum is 1/6. The next term is −1/8. The next two terms are 1/5 and −1/10, whose sum is 1/10. In general, since every odd integer occurs once positively and every even integers occur once negatively (half of them as multiples of 4, the other half as twice odd integers), the sum is composed of blocks of three which can be simplified as:

 


Hence, the above series can in fact be written as:

 

which is half the sum originally, and can only equate to the original sequence if the value were zero. This series can be demonstrated to be greater than zero by the proof of Leibniz's theorem using that the second partial sum is half.[7] Alternatively, the value of   which it converges to, cannot be zero. Hence, the value of the sequence is shown to depend on the order in which series is computed.

It is true that the sequence:

 

contains all elements in the sequence:

 


However, since the summation is defined as   and  , the order of the terms can influence the limit.[7]

Getting an arbitrary sum edit

An efficient way to recover and generalize the result of the previous section is to use the fact that

 

where γ is the Euler–Mascheroni constant, and where the notation o(1) denotes a quantity that depends upon the current variable (here, the variable is n) in such a way that this quantity goes to 0 when the variable tends to infinity.

It follows that the sum of q even terms satisfies

 

and by taking the difference, one sees that the sum of p odd terms satisfies

 

Suppose that two positive integers a and b are given, and that a rearrangement of the alternating harmonic series is formed by taking, in order, a positive terms from the alternating harmonic series, followed by b negative terms, and repeating this pattern at infinity (the alternating series itself corresponds to a = b = 1, the example in the preceding section corresponds to a = 1, b = 2):

 

Then the partial sum of order (a + b)n of this rearranged series contains p = an positive odd terms and q = bn negative even terms, hence

 

It follows that the sum of this rearranged series is[8]

 

Suppose now that, more generally, a rearranged series of the alternating harmonic series is organized in such a way that the ratio pn/qn between the number of positive and negative terms in the partial sum of order n tends to a positive limit r. Then, the sum of such a rearrangement will be

 

and this explains that any real number x can be obtained as sum of a rearranged series of the alternating harmonic series: it suffices to form a rearrangement for which the limit r is equal to e2x/ 4.

Proof edit

Existence of a rearrangement that sums to any positive real M edit

Riemann's description of the theorem and its proof reads in full:[9]

… infinite series fall into two distinct classes, depending on whether or not they remain convergent when all the terms are made positive. In the first class the terms can be arbitrarily rearranged; in the second, on the other hand, the value is dependent on the ordering of the terms. Indeed, if we denote the positive terms of a series in the second class by a1, a2, a3, ... and the negative terms by b1, −b2, −b3, ... then it is clear that Σa as well as Σb must be infinite. For if they were both finite, the series would still be convergent after making all the signs the same. If only one were infinite, then the series would diverge. Clearly now an arbitrarily given value C can be obtained by a suitable reordering of the terms. We take alternately the positive terms of the series until the sum is greater than C, and then the negative terms until the sum is less than C. The deviation from C never amounts to more than the size of the term at the last place the signs were switched. Now, since the number a as well as the numbers b become infinitely small with increasing index, so also are the deviations from C. If we proceed sufficiently far in the series, the deviation becomes arbitrarily small, that is, the series converges to C.

This can be given more detail as follows.[10] Recall that a conditionally convergent series of real terms has both infinitely many negative terms and infinitely many positive terms. First, define two quantities,   and   by:

 

That is, the series   includes all an positive, with all negative terms replaced by zeroes, and the series   includes all an negative, with all positive terms replaced by zeroes. Since   is conditionally convergent, both the 'positive' and the 'negative' series diverge. Let M be any real number. Take just enough of the positive terms   so that their sum exceeds M. That is, let p1 be the smallest positive integer such that

 

This is possible because the partial sums of the   series tend to  . Now let q1 be the smallest positive integer such that

 

This number exists because the partial sums of   tend to  . Now continue inductively, defining p2 as the smallest integer larger than p1 such that

 

and so on. The result may be viewed as a new sequence

 

Furthermore the partial sums of this new sequence converge to M. This can be seen from the fact that for any i,

 

with the first inequality holding due to the fact that pi+1 has been defined as the smallest number larger than pi which makes the second inequality true; as a consequence, it holds that

 

Since the right-hand side converges to zero due to the assumption of conditional convergence, this shows that the (pi+1 + qi)'th partial sum of the new sequence converges to M as i increases. Similarly, the (pi+1 + qi+1)'th partial sum also converges to M. Since the (pi+1 + qi + 1)'th, (pi+1 + qi + 2)'th, ... (pi+1 + qi+1 − 1)'th partial sums are valued between the (pi+1 + qi)'th and (pi+1 + qi+1)'th partial sums, it follows that the whole sequence of partial sums converges to M.

Every entry in the original sequence an appears in this new sequence whose partial sums converge to M. Those entries of the original sequence which are zero will appear twice in the new sequence (once in the 'positive' sequence and once in the 'negative' sequence), and every second such appearance can be removed, which does not affect the summation in any way. The new sequence is thus a permutation of the original sequence.

Existence of a rearrangement that diverges to infinity edit

Let   be a conditionally convergent series. The following is a proof that there exists a rearrangement of this series that tends to   (a similar argument can be used to show that   can also be attained).

The above proof of Riemann's original formulation only needs to be modified so that pi+1 is selected as the smallest integer larger than pi such that

 

and with qi+1 selected as the smallest integer larger than qi such that

 

The choice of i+1 on the left-hand sides is immaterial, as it could be replaced by any sequence increasing to infinity. Since   converges to zero as n increases, for sufficiently large i there is

 

and this proves (just as with the analysis of convergence above) that the sequence of partial sums of the new sequence diverge to infinity.

Existence of a rearrangement that fails to approach any limit, finite or infinite edit

The above proof only needs to be modified so that pi+1 is selected as the smallest integer larger than pi such that

 

and with qi+1 selected as the smallest integer larger than qi such that

 

This directly shows that the sequence of partial sums contains infinitely many entries which are larger than 1, and also infinitely many entries which are less than −1, so that the sequence of partial sums cannot converge.

Generalizations edit

Sierpiński theorem edit

Given an infinite series  , we may consider a set of "fixed points"  , and study the real numbers that the series can sum to if we are only allowed to permute indices in  . That is, we let

 
With this notation, we have:
  • If   is finite, then  . Here   means symmetric difference.
  • If   then  .
  • If the series is an absolutely convergent sum, then   for any  .
  • If the series is a conditionally convergent sum, then by Riemann series theorem,  .

Sierpiński proved that rearranging only the positive terms one can obtain a series converging to any prescribed value less than or equal to the sum of the original series, but larger values in general can not be attained.[11][12][13] That is, let   be a conditionally convergent sum, then   contains  , but there is no guarantee that it contains any other number.

More generally, let   be an ideal of  , then we can define  .

Let   be the set of all asymptotic density zero sets  , that is,  . It's clear that   is an ideal of  .

(Władysław, 2007)[14] —  If   is a conditionally convergent sum, then   (that is, it is sufficient to rearrange a set of indices of asymptotic density zero).

Proof sketch: Given  , a conditionally convergent sum, construct some   such that   and   are both conditionally convergent. Then, rearranging   suffices to converge to any number in  .

Filipów and Szuca proved that other ideals also have this property.[15]

Steinitz's theorem edit

Given a converging series  of complex numbers, several cases can occur when considering the set of possible sums for all series   obtained by rearranging (permuting) the terms of that series:

  • the series   may converge unconditionally; then, all rearranged series converge, and have the same sum: the set of sums of the rearranged series reduces to one point;
  • the series   may fail to converge unconditionally; if S denotes the set of sums of those rearranged series that converge, then, either the set S is a line L in the complex plane C, of the form
     
    or the set S is the whole complex plane C.

More generally, given a converging series of vectors in a finite-dimensional real vector space E, the set of sums of converging rearranged series is an affine subspace of E.

See also edit

References edit

  1. ^ Cauchy 1833, Section 8; Apostol 1967, p. 411.
  2. ^ Dirichlet 1837, Section 1.
  3. ^ Riemann 1868.
  4. ^ Kline 1990, p. 966.
  5. ^ Apostol 1967, Section 10.21; Apostol 1974, Section 8.18; Rudin 1976, Theorem 3.54; Whittaker & Watson 2021, Section II.17.
  6. ^ Banaszczyk 1991, Section 10; Mauldin 2015, Problem 28 and Problem 106.
  7. ^ a b Spivak, Michael (2008). Calculus (4th ed.). Houston, Texas: Publish or Perish. pp. 482–483. ISBN 978-0-914098-91-1.
  8. ^ Apostol, Tom M. (1991-01-16). Calculus, Volume 1. John Wiley & Sons. p. 416. ISBN 978-0-471-00005-1.
  9. ^ Riemann 1868, p. 97, quoted from the 2004 English translation.
  10. ^ Apostol 1967, Section 10.21; Whittaker & Watson 2021, Section II.17.
  11. ^ Sierpiński, Wacław (1910). "Przyczynek do teoryi szeregów rozbieżnych [Contribution à la théorie des séries divergentes]" [Contribution to the theory of divergent series]. Sprawozdania Z Posiedzen Towarzystwa Naukowego Warszawskiego (in Polish). 3: 89–93.
  12. ^ Sierpiński, Wacław (1910). "Uwaga do twierdzenia Riemanna o szeregach warunkowo zbieżnych [Remarque sur le théorème de Riemann relatif aux séries semiconvergentes]" [Remark on Riemann's theorem relating to semi-convergent series]. Prace Matematyczno-Fizyczne (in Polish). 21 (1): 17–20.
  13. ^ Sierpiński, Wacław (1911). "Sur une propriété des séries qui ne sont pas absolument convergentes [O pewnej własności szeregów warunkowo zbieżnych]". Bulletin International de l'Académie des Sciences de Cracovie, Séries A: 149–158.
  14. ^ Wilczyński, Władysław (2007). "On Riemann derangement theorem". Słupskie Prace Matematyczno-Fizyczne. 4: 79–82.
  15. ^ Filipów, Rafał; Szuca, Piotr (February 2010). "Rearrangement of conditionally convergent series on a small set". Journal of Mathematical Analysis and Applications. 362 (1): 64–71. doi:10.1016/j.jmaa.2009.07.029.

External links edit

riemann, series, theorem, mathematics, also, called, riemann, rearrangement, theorem, named, after, 19th, century, german, mathematician, bernhard, riemann, says, that, infinite, series, real, numbers, conditionally, convergent, then, terms, arranged, permutat. In mathematics the Riemann series theorem also called the Riemann rearrangement theorem named after 19th century German mathematician Bernhard Riemann says that if an infinite series of real numbers is conditionally convergent then its terms can be arranged in a permutation so that the new series converges to an arbitrary real number or diverges This implies that a series of real numbers is absolutely convergent if and only if it is unconditionally convergent As an example the series 1 1 1 2 1 2 1 3 1 3 converges to 0 for a sufficiently large number of terms the partial sum gets arbitrarily near to 0 but replacing all terms with their absolute values gives 1 1 1 2 1 2 1 3 1 3 which sums to infinity Thus the original series is conditionally convergent and can be rearranged by taking the first two positive terms followed by the first negative term followed by the next two positive terms and then the next negative term etc to give a series that converges to a different sum 1 1 2 1 1 3 1 4 1 2 ln 2 More generally using this procedure with p positives followed by q negatives gives the sum ln p q Other rearrangements give other finite sums or do not converge to any sum Contents 1 History 2 Definitions 3 Statement of the theorem 4 Alternating harmonic series 4 1 Changing the sum 4 2 Getting an arbitrary sum 5 Proof 5 1 Existence of a rearrangement that sums to any positive real M 5 2 Existence of a rearrangement that diverges to infinity 5 3 Existence of a rearrangement that fails to approach any limit finite or infinite 6 Generalizations 6 1 Sierpinski theorem 6 2 Steinitz s theorem 7 See also 8 References 9 External linksHistory editIt is a basic result that the sum of finitely many numbers does not depend on the order in which they are added For example 2 6 7 7 2 6 The observation that the sum of an infinite sequence of numbers can depend on the ordering of the summands is commonly attributed to Augustin Louis Cauchy in 1833 1 He analyzed the alternating harmonic series showing that certain rearrangements of its summands result in different limits Around the same time Peter Gustav Lejeune Dirichlet highlighted that such phenomena are ruled out in the context of absolute convergence and gave further examples of Cauchy s phenomenon for some other series which fail to be absolutely convergent 2 In the course of his analysis of Fourier series and the theory of Riemann integration Bernhard Riemann gave a full characterization of the rearrangement phenomena 3 He proved that in the case of a convergent series which does not converge absolutely known as conditional convergence rearrangements can be found so that the new series converges to any arbitrarily prescribed real number 4 Riemann s theorem is now considered as a basic part of the field of mathematical analysis 5 For any series one may consider the set of all possible sums corresponding to all possible rearrangements of the summands Riemann s theorem can be formulated as saying that for a series of real numbers this set is either empty a single point in the case of absolute convergence or the entire real number line in the case of conditional convergence In this formulation Riemann s theorem was extended by Paul Levy and Ernst Steinitz to series whose summands are complex numbers or even more generally elements of a finite dimensional real vector space They proved that the set of possible sums forms a real affine subspace Extensions of the Levy Steinitz theorem to series in infinite dimensional spaces have been considered by a number of authors 6 Definitions editA series n 1 a n textstyle sum n 1 infty a n nbsp converges if there exists a value ℓ displaystyle ell nbsp such that the sequence of the partial sums S 1 S 2 S 3 S n k 1 n a k displaystyle S 1 S 2 S 3 ldots quad S n sum k 1 n a k nbsp converges to ℓ displaystyle ell nbsp That is for any e gt 0 there exists an integer N such that if n N then S n ℓ e displaystyle left vert S n ell right vert leq varepsilon nbsp A series converges conditionally if the series n 1 a n textstyle sum n 1 infty a n nbsp converges but the series n 1 a n textstyle sum n 1 infty left vert a n right vert nbsp diverges A permutation is simply a bijection from the set of positive integers to itself This means that if s displaystyle sigma nbsp is a permutation then for any positive integer b displaystyle b nbsp there exists exactly one positive integer a displaystyle a nbsp such that s a b displaystyle sigma a b nbsp In particular if x y displaystyle x neq y nbsp then s x s y displaystyle sigma x neq sigma y nbsp Statement of the theorem editSuppose that a 1 a 2 a 3 displaystyle a 1 a 2 a 3 ldots nbsp is a sequence of real numbers and that n 1 a n textstyle sum n 1 infty a n nbsp is conditionally convergent Let M displaystyle M nbsp be a real number Then there exists a permutation s displaystyle sigma nbsp such that n 1 a s n M displaystyle sum n 1 infty a sigma n M nbsp There also exists a permutation s displaystyle sigma nbsp such that n 1 a s n displaystyle sum n 1 infty a sigma n infty nbsp The sum can also be rearranged to diverge to displaystyle infty nbsp or to fail to approach any limit finite or infinite Alternating harmonic series editChanging the sum edit The alternating harmonic series is a classic example of a conditionally convergent series n 1 1 n 1 n displaystyle sum n 1 infty frac 1 n 1 n nbsp is convergent whereas n 1 1 n 1 n n 1 1 n displaystyle sum n 1 infty left frac 1 n 1 n right sum n 1 infty frac 1 n nbsp is the ordinary harmonic series which diverges Although in standard presentation the alternating harmonic series converges to ln 2 its terms can be arranged to converge to any number or even to diverge One instance of this is as follows Begin with the series written in the usual order ln 2 1 1 2 1 3 1 4 1 5 1 6 1 7 1 8 1 9 displaystyle ln 2 1 frac 1 2 frac 1 3 frac 1 4 frac 1 5 frac 1 6 frac 1 7 frac 1 8 frac 1 9 cdots nbsp and rearrange and regroup the terms as 1 1 2 1 4 1 3 1 6 1 8 1 5 1 10 1 12 1 1 2 1 4 1 3 1 6 1 8 1 5 1 10 1 12 displaystyle begin aligned amp 1 frac 1 2 frac 1 4 frac 1 3 frac 1 6 frac 1 8 frac 1 5 frac 1 10 frac 1 12 cdots amp left 1 frac 1 2 right frac 1 4 left frac 1 3 frac 1 6 right frac 1 8 left frac 1 5 frac 1 10 right frac 1 12 cdots end aligned nbsp where the pattern is the first two terms are 1 and 1 2 whose sum is 1 2 The next term is 1 4 The next two terms are 1 3 and 1 6 whose sum is 1 6 The next term is 1 8 The next two terms are 1 5 and 1 10 whose sum is 1 10 In general since every odd integer occurs once positively and every even integers occur once negatively half of them as multiples of 4 the other half as twice odd integers the sum is composed of blocks of three which can be simplified as 1 2 k 1 1 2 2 k 1 1 4 k 1 2 2 k 1 1 4 k k 1 2 displaystyle left frac 1 2k 1 frac 1 2 2k 1 right frac 1 4k left frac 1 2 2k 1 right frac 1 4k quad k 1 2 dots nbsp Hence the above series can in fact be written as 1 2 1 4 1 6 1 8 1 10 1 2 2 k 1 1 2 2 k 1 2 1 1 2 1 3 1 2 ln 2 displaystyle begin aligned amp frac 1 2 frac 1 4 frac 1 6 frac 1 8 frac 1 10 cdots frac 1 2 2k 1 frac 1 2 2k cdots amp frac 1 2 left 1 frac 1 2 frac 1 3 cdots right frac 1 2 ln 2 end aligned nbsp which is half the sum originally and can only equate to the original sequence if the value were zero This series can be demonstrated to be greater than zero by the proof of Leibniz s theorem using that the second partial sum is half 7 Alternatively the value of ln 2 displaystyle ln 2 nbsp which it converges to cannot be zero Hence the value of the sequence is shown to depend on the order in which series is computed It is true that the sequence b n 1 1 2 1 4 1 3 1 6 1 8 1 5 1 10 1 12 1 7 1 14 1 16 displaystyle b n 1 frac 1 2 frac 1 4 frac 1 3 frac 1 6 frac 1 8 frac 1 5 frac 1 10 frac 1 12 frac 1 7 frac 1 14 frac 1 16 cdots nbsp contains all elements in the sequence a n 1 1 2 1 3 1 4 1 5 1 6 1 7 1 8 1 9 1 10 1 11 1 12 1 13 1 14 1 15 displaystyle a n 1 frac 1 2 frac 1 3 frac 1 4 frac 1 5 frac 1 6 frac 1 7 frac 1 8 frac 1 9 frac 1 10 frac 1 11 frac 1 12 frac 1 13 frac 1 14 frac 1 15 cdots nbsp However since the summation is defined as n 1 a n lim n a 1 a 2 a n displaystyle sum n 1 infty a n lim n to infty left a 1 a 2 cdots a n right nbsp and n 1 b n lim n b 1 b 2 b n displaystyle sum n 1 infty b n lim n to infty left b 1 b 2 cdots b n right nbsp the order of the terms can influence the limit 7 Getting an arbitrary sum edit An efficient way to recover and generalize the result of the previous section is to use the fact that 1 1 2 1 3 1 n g ln n o 1 displaystyle 1 1 over 2 1 over 3 cdots 1 over n gamma ln n o 1 nbsp where g is the Euler Mascheroni constant and where the notation o 1 denotes a quantity that depends upon the current variable here the variable is n in such a way that this quantity goes to 0 when the variable tends to infinity It follows that the sum of q even terms satisfies 1 2 1 4 1 6 1 2 q 1 2 g 1 2 ln q o 1 displaystyle 1 over 2 1 over 4 1 over 6 cdots 1 over 2q 1 over 2 gamma 1 over 2 ln q o 1 nbsp and by taking the difference one sees that the sum of p odd terms satisfies 1 1 3 1 5 1 2 p 1 1 2 g 1 2 ln p ln 2 o 1 displaystyle 1 1 over 3 1 over 5 cdots 1 over 2p 1 1 over 2 gamma 1 over 2 ln p ln 2 o 1 nbsp Suppose that two positive integers a and b are given and that a rearrangement of the alternating harmonic series is formed by taking in order a positive terms from the alternating harmonic series followed by b negative terms and repeating this pattern at infinity the alternating series itself corresponds to a b 1 the example in the preceding section corresponds to a 1 b 2 1 1 3 1 2 a 1 1 2 1 4 1 2 b 1 2 a 1 1 4 a 1 1 2 b 2 displaystyle 1 1 over 3 cdots 1 over 2a 1 1 over 2 1 over 4 cdots 1 over 2b 1 over 2a 1 cdots 1 over 4a 1 1 over 2b 2 cdots nbsp Then the partial sum of order a b n of this rearranged series contains p an positive odd terms and q bn negative even terms hence S a b n 1 2 ln p ln 2 1 2 ln q o 1 1 2 ln a b ln 2 o 1 displaystyle S a b n 1 over 2 ln p ln 2 1 over 2 ln q o 1 1 over 2 ln left frac a b right ln 2 o 1 nbsp It follows that the sum of this rearranged series is 8 1 2 ln a b ln 2 ln 2 a b displaystyle 1 over 2 ln left frac a b right ln 2 ln left 2 sqrt frac a b right nbsp Suppose now that more generally a rearranged series of the alternating harmonic series is organized in such a way that the ratio pn qn between the number of positive and negative terms in the partial sum of order n tends to a positive limit r Then the sum of such a rearrangement will be ln 2 r displaystyle ln left 2 sqrt r right nbsp and this explains that any real number x can be obtained as sum of a rearranged series of the alternating harmonic series it suffices to form a rearrangement for which the limit r is equal to e2x 4 Proof editExistence of a rearrangement that sums to any positive real M edit Riemann s description of the theorem and its proof reads in full 9 infinite series fall into two distinct classes depending on whether or not they remain convergent when all the terms are made positive In the first class the terms can be arbitrarily rearranged in the second on the other hand the value is dependent on the ordering of the terms Indeed if we denote the positive terms of a series in the second class by a1 a2 a3 and the negative terms by b1 b2 b3 then it is clear that Sa as well as Sb must be infinite For if they were both finite the series would still be convergent after making all the signs the same If only one were infinite then the series would diverge Clearly now an arbitrarily given value C can be obtained by a suitable reordering of the terms We take alternately the positive terms of the series until the sum is greater than C and then the negative terms until the sum is less than C The deviation from C never amounts to more than the size of the term at the last place the signs were switched Now since the number a as well as the numbers b become infinitely small with increasing index so also are the deviations from C If we proceed sufficiently far in the series the deviation becomes arbitrarily small that is the series converges to C This can be given more detail as follows 10 Recall that a conditionally convergent series of real terms has both infinitely many negative terms and infinitely many positive terms First define two quantities a n displaystyle a n nbsp and a n displaystyle a n nbsp by a n a n if a n 0 0 if a n lt 0 a n 0 if a n 0 a n if a n lt 0 displaystyle a n begin cases a n amp text if a n geq 0 0 amp text if a n lt 0 end cases qquad a n begin cases 0 amp text if a n geq 0 a n amp text if a n lt 0 end cases nbsp That is the series n 1 a n textstyle sum n 1 infty a n nbsp includes all an positive with all negative terms replaced by zeroes and the series n 1 a n textstyle sum n 1 infty a n nbsp includes all an negative with all positive terms replaced by zeroes Since n 1 a n textstyle sum n 1 infty a n nbsp is conditionally convergent both the positive and the negative series diverge Let M be any real number Take just enough of the positive terms a n displaystyle a n nbsp so that their sum exceeds M That is let p1 be the smallest positive integer such that M lt n 1 p 1 a n displaystyle M lt sum n 1 p 1 a n nbsp This is possible because the partial sums of the a n displaystyle a n nbsp series tend to displaystyle infty nbsp Now let q1 be the smallest positive integer such that M gt n 1 p 1 a n n 1 q 1 a n displaystyle M gt sum n 1 p 1 a n sum n 1 q 1 a n nbsp This number exists because the partial sums of a n displaystyle a n nbsp tend to displaystyle infty nbsp Now continue inductively defining p2 as the smallest integer larger than p1 such that M lt n 1 p 2 a n n 1 q 1 a n displaystyle M lt sum n 1 p 2 a n sum n 1 q 1 a n nbsp and so on The result may be viewed as a new sequence a 1 a p 1 a 1 a q 1 a p 1 1 a p 2 a q 1 1 a q 2 a p 2 1 displaystyle a 1 ldots a p 1 a 1 ldots a q 1 a p 1 1 ldots a p 2 a q 1 1 ldots a q 2 a p 2 1 ldots nbsp Furthermore the partial sums of this new sequence converge to M This can be seen from the fact that for any i n 1 p i 1 1 a n n 1 q i a n M lt n 1 p i 1 a n n 1 q i a n displaystyle sum n 1 p i 1 1 a n sum n 1 q i a n leq M lt sum n 1 p i 1 a n sum n 1 q i a n nbsp with the first inequality holding due to the fact that pi 1 has been defined as the smallest number larger than pi which makes the second inequality true as a consequence it holds that 0 lt n 1 p i 1 a n n 1 q i a n M a p i 1 displaystyle 0 lt left sum n 1 p i 1 a n sum n 1 q i a n right M leq a p i 1 nbsp Since the right hand side converges to zero due to the assumption of conditional convergence this shows that the pi 1 qi th partial sum of the new sequence converges to M as i increases Similarly the pi 1 qi 1 th partial sum also converges to M Since the pi 1 qi 1 th pi 1 qi 2 th pi 1 qi 1 1 th partial sums are valued between the pi 1 qi th and pi 1 qi 1 th partial sums it follows that the whole sequence of partial sums converges to M Every entry in the original sequence an appears in this new sequence whose partial sums converge to M Those entries of the original sequence which are zero will appear twice in the new sequence once in the positive sequence and once in the negative sequence and every second such appearance can be removed which does not affect the summation in any way The new sequence is thus a permutation of the original sequence Existence of a rearrangement that diverges to infinity edit Let i 1 a i textstyle sum i 1 infty a i nbsp be a conditionally convergent series The following is a proof that there exists a rearrangement of this series that tends to displaystyle infty nbsp a similar argument can be used to show that displaystyle infty nbsp can also be attained The above proof of Riemann s original formulation only needs to be modified so that pi 1 is selected as the smallest integer larger than pi such that i 1 lt n 1 p i 1 a n n 1 q i a n displaystyle i 1 lt sum n 1 p i 1 a n sum n 1 q i a n nbsp and with qi 1 selected as the smallest integer larger than qi such that i 1 gt n 1 p i 1 a n n 1 q i 1 a n displaystyle i 1 gt sum n 1 p i 1 a n sum n 1 q i 1 a n nbsp The choice of i 1 on the left hand sides is immaterial as it could be replaced by any sequence increasing to infinity Since a n displaystyle a n nbsp converges to zero as n increases for sufficiently large i there is n 1 p i 1 a n n 1 q i 1 a n gt i displaystyle sum n 1 p i 1 a n sum n 1 q i 1 a n gt i nbsp and this proves just as with the analysis of convergence above that the sequence of partial sums of the new sequence diverge to infinity Existence of a rearrangement that fails to approach any limit finite or infinite edit The above proof only needs to be modified so that pi 1 is selected as the smallest integer larger than pi such that 1 lt n 1 p i 1 a n n 1 q i a n displaystyle 1 lt sum n 1 p i 1 a n sum n 1 q i a n nbsp and with qi 1 selected as the smallest integer larger than qi such that 1 gt n 1 p i 1 a n n 1 q i 1 a n displaystyle 1 gt sum n 1 p i 1 a n sum n 1 q i 1 a n nbsp This directly shows that the sequence of partial sums contains infinitely many entries which are larger than 1 and also infinitely many entries which are less than 1 so that the sequence of partial sums cannot converge Generalizations editSierpinski theorem edit Given an infinite series a a 1 a 2 displaystyle a a 1 a 2 nbsp we may consider a set of fixed points I N displaystyle I subset mathbb N nbsp and study the real numbers that the series can sum to if we are only allowed to permute indices in I displaystyle I nbsp That is we letS a I n N a p n p is a permutation on N such that n I p n n and the summation converges displaystyle S a I left sum n in mathbb N a pi n pi text is a permutation on mathbb N text such that forall n not in I pi n n text and the summation converges right nbsp With this notation we have If I I displaystyle I mathbin triangle I nbsp is finite then S a I S a I displaystyle S a I S a I nbsp Here displaystyle triangle nbsp means symmetric difference If I I displaystyle I subset I nbsp then S a I S a I displaystyle S a I subset S a I nbsp If the series is an absolutely convergent sum then S a I n N a n displaystyle S a I left sum n in mathbb N a n right nbsp for any I displaystyle I nbsp If the series is a conditionally convergent sum then by Riemann series theorem S a N displaystyle S a mathbb N infty infty nbsp Sierpinski proved that rearranging only the positive terms one can obtain a series converging to any prescribed value less than or equal to the sum of the original series but larger values in general can not be attained 11 12 13 That is let a displaystyle a nbsp be a conditionally convergent sum then S a n N a n gt 0 displaystyle S a n in mathbb N a n gt 0 nbsp contains n N a n displaystyle left infty sum n in mathbb N a n right nbsp but there is no guarantee that it contains any other number More generally let J displaystyle J nbsp be an ideal of N displaystyle mathbb N nbsp then we can define S a J I J S a I displaystyle S a J cup I in J S a I nbsp Let J d displaystyle J d nbsp be the set of all asymptotic density zero sets I N displaystyle I subset mathbb N nbsp that is lim n 0 n I n 0 displaystyle lim n to infty frac 0 n cap I n 0 nbsp It s clear that J d displaystyle J d nbsp is an ideal of N displaystyle mathbb N nbsp Wladyslaw 2007 14 If a displaystyle a nbsp is a conditionally convergent sum then S a J d displaystyle S a J d infty infty nbsp that is it is sufficient to rearrange a set of indices of asymptotic density zero Proof sketch Given a displaystyle a nbsp a conditionally convergent sum construct some I J d displaystyle I in J d nbsp such that n I a n displaystyle sum n in I a n nbsp and n I a n displaystyle sum n not in I a n nbsp are both conditionally convergent Then rearranging n I a n displaystyle sum n in I a n nbsp suffices to converge to any number in displaystyle infty infty nbsp Filipow and Szuca proved that other ideals also have this property 15 Steinitz s theorem edit Main article Levy Steinitz theorem Given a converging series a n textstyle sum a n nbsp of complex numbers several cases can occur when considering the set of possible sums for all series a s n textstyle sum a sigma n nbsp obtained by rearranging permuting the terms of that series the series a n textstyle sum a n nbsp may converge unconditionally then all rearranged series converge and have the same sum the set of sums of the rearranged series reduces to one point the series a n textstyle sum a n nbsp may fail to converge unconditionally if S denotes the set of sums of those rearranged series that converge then either the set S is a line L in the complex plane C of the form L a t b t R a b C b 0 displaystyle L a tb t in mathbb R quad a b in mathbb C b neq 0 nbsp or the set S is the whole complex plane C More generally given a converging series of vectors in a finite dimensional real vector space E the set of sums of converging rearranged series is an affine subspace of E See also editAbsolute convergence Rearrangements and unconditional convergenceReferences edit Cauchy 1833 Section 8 Apostol 1967 p 411 Dirichlet 1837 Section 1 Riemann 1868 Kline 1990 p 966 Apostol 1967 Section 10 21 Apostol 1974 Section 8 18 Rudin 1976 Theorem 3 54 Whittaker amp Watson 2021 Section II 17 Banaszczyk 1991 Section 10 Mauldin 2015 Problem 28 and Problem 106 a b Spivak Michael 2008 Calculus 4th ed Houston Texas Publish or Perish pp 482 483 ISBN 978 0 914098 91 1 Apostol Tom M 1991 01 16 Calculus Volume 1 John Wiley amp Sons p 416 ISBN 978 0 471 00005 1 Riemann 1868 p 97 quoted from the 2004 English translation Apostol 1967 Section 10 21 Whittaker amp Watson 2021 Section II 17 Sierpinski Waclaw 1910 Przyczynek do teoryi szeregow rozbieznych Contribution a la theorie des series divergentes Contribution to the theory of divergent series Sprawozdania Z Posiedzen Towarzystwa Naukowego Warszawskiego in Polish 3 89 93 Sierpinski Waclaw 1910 Uwaga do twierdzenia Riemanna o szeregach warunkowo zbieznych Remarque sur le theoreme de Riemann relatif aux series semiconvergentes Remark on Riemann s theorem relating to semi convergent series Prace Matematyczno Fizyczne in Polish 21 1 17 20 Sierpinski Waclaw 1911 Sur une propriete des series qui ne sont pas absolument convergentes O pewnej wlasnosci szeregow warunkowo zbieznych Bulletin International de l Academie des Sciences de Cracovie Series A 149 158 Wilczynski Wladyslaw 2007 On Riemann derangement theorem Slupskie Prace Matematyczno Fizyczne 4 79 82 Filipow Rafal Szuca Piotr February 2010 Rearrangement of conditionally convergent series on a small set Journal of Mathematical Analysis and Applications 362 1 64 71 doi 10 1016 j jmaa 2009 07 029 Apostol Tom M 1967 Calculus Volume I One variable calculus with an introduction to linear algebra Second edition of 1961 original ed New York John Wiley amp Sons Inc ISBN 0 471 00005 1 MR 0214705 Zbl 0148 28201 Apostol Tom M 1974 Mathematical analysis Second edition of 1957 original ed Reading MA Addison Wesley Publishing Co MR 0344384 Zbl 0309 26002 Banaszczyk Wojciech 1991 Additive subgroups of topological vector spaces Lecture Notes in Mathematics Vol 1466 Berlin Springer Verlag pp 93 109 doi 10 1007 BFb0089147 ISBN 3 540 53917 4 MR 1119302 Zbl 0743 46002 Cauchy M Augustin Louis 1833 Resumes analytiques Turin L imprimerie royale Dirichlet P G L 1837 Beweis des Satzes dass jede unbegrenzte arithmetische Progression deren erstes Glied und Differenz ganze Zahlen ohne gemeinschaftlichen Factor sind unendlich viele Primzahlen enthalt Abhandlungen der Koniglich Preussischen Akademie der Wissenschaften 45 81 Lejeune Dirichlet G 1889 Beweis des Satzes dass jede unbegrenzte arithmetische Progression deren erstes Glied und Differenz ganze Zahlen ohne gemeinschaftlichen Factor sind unendlich viele Primzahlen enthalt In Kronecker L ed Werke Band I Berlin Dietrich Reimer Verlag pp 313 342 JFM 21 0016 01 MR 0249268 Kline Morris 1990 Mathematical thought from ancient to modern times Volume 3 Second edition of 1972 original ed New York The Clarendon Press ISBN 0 19 506137 3 MR 1058203 Zbl 0864 01001 Mauldin R Daniel ed 2015 The Scottish Book Mathematics from the Scottish Cafe with selected problems from the new Scottish Book Including selected papers presented at the Scottish Book Conference held at North Texas University Denton TX May 1979 Second edition of 1981 original ed Springer Cham doi 10 1007 978 3 319 22897 6 ISBN 978 3 319 22896 9 MR 3242261 Zbl 1331 01039 Riemann Bernhard 1868 Uber die Darstellbarkeit einer Function durch eine trigonometrische Reihe Abhandlungen der Koniglichen Gesellschaft der Wissenschaften zu Gottingen 13 87 132 JFM 01 0131 03 Riemann Bernhard 2004 On the representation of a function by a trigonometric series Collected Papers Translated by Baker Roger Christenson Charles Orde Henry Translation of 1892 German edition Heber City UT Kendrick Press ISBN 0 9740427 2 2 MR 2121437 Zbl 1101 01013 Rudin Walter 1976 Principles of mathematical analysis International Series in Pure and Applied Mathematics Third edition of 1953 original ed New York McGraw Hill Book Co MR 0385023 Zbl 0346 26002 Whittaker E T Watson G N 2021 Moll Victor H ed A course of modern analysis an introduction to the general theory of infinite processes and of analytic functions with an account of the principal transcendental functions With a foreword by S J Patterson Fifth edition of 1902 original ed Cambridge Cambridge University Press doi 10 1017 9781009004091 ISBN 978 1 316 51893 9 MR 4286926 Zbl 1468 30001 External links editWeisstein Eric W Riemann Series Theorem MathWorld Retrieved February 1 2023 Retrieved from https en wikipedia org w index php title Riemann series theorem amp oldid 1211198752, 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.