Open main menu
Home
Random
Recent changes
Special pages
Community portal
Preferences
About Wikipedia
Disclaimers
Incubator escapee wiki
Search
User menu
Talk
Dark mode
Contributions
Create account
Log in
Editing
Basel problem
(section)
Warning:
You are not logged in. Your IP address will be publicly visible if you make any edits. If you
log in
or
create an account
, your edits will be attributed to your username, along with other benefits.
Anti-spam check. Do
not
fill this in!
===The proof=== [[File:limit circle FbN.jpeg|thumb|The inequality<br> <math>\tfrac{1}{2}r^2\tan\theta > \tfrac{1}{2}r^2\theta > \tfrac{1}{2}r^2\sin\theta</math><br> is shown pictorially for any <math>\theta \in (0, \pi/2)</math>. The three terms are the areas of the triangle OAC, circle section OAB, and the triangle OAB. Taking reciprocals and squaring gives<br> <math>\cot^2\theta<\tfrac{1}{\theta^2}<\csc^2\theta</math>.]] The main idea behind the proof is to bound the partial (finite) sums <math display=block>\sum_{k=1}^m \frac{1}{k^2} = \frac{1}{1^2} + \frac{1}{2^2} + \cdots + \frac{1}{m^2}</math> between two expressions, each of which will tend to {{sfrac|{{pi}}<sup>2</sup>|6}} as {{math|''m''}} approaches infinity. The two expressions are derived from identities involving the [[cotangent]] and [[cosecant]] functions. These identities are in turn derived from [[de Moivre's formula]], and we now turn to establishing these identities. Let {{math|''x''}} be a real number with {{math|0 < ''x'' < {{sfrac|{{pi}}|2}}}}, and let {{math|''n''}} be a positive odd integer. Then from de Moivre's formula and the definition of the cotangent function, we have <math display=block>\begin{align} \frac{\cos (nx) + i \sin (nx)}{\sin^n x} &= \frac{(\cos x + i\sin x)^n}{\sin^n x} \\[4pt] &= \left(\frac{\cos x + i \sin x}{\sin x}\right)^n \\[4pt] &= (\cot x + i)^n. \end{align}</math> From the [[binomial theorem]], we have <math display=block>\begin{align} (\cot x + i)^n = & {n \choose 0} \cot^n x + {n \choose 1} (\cot^{n - 1} x)i + \cdots + {n \choose {n - 1}} (\cot x)i^{n - 1} + {n \choose n} i^n \\[6pt] = & \Bigg( {n \choose 0} \cot^n x - {n \choose 2} \cot^{n - 2} x \pm \cdots \Bigg) \; + \; i\Bigg( {n \choose 1} \cot^{n-1} x - {n \choose 3} \cot^{n - 3} x \pm \cdots \Bigg). \end{align}</math> Combining the two equations and equating imaginary parts gives the identity <math display=block>\frac{\sin (nx)}{\sin^n x} = \Bigg( {n \choose 1} \cot^{n - 1} x - {n \choose 3} \cot^{n - 3} x \pm \cdots \Bigg).</math> We take this identity, fix a positive integer {{math|''m''}}, set {{math|''n'' {{=}} 2''m'' + 1}}, and consider {{math|''x<sub>r</sub>'' {{=}} {{sfrac|''r''{{pi}}|2''m'' + 1}}}} for {{math|''r'' {{=}} 1, 2, ..., ''m''}}. Then {{math|''nx<sub>r</sub>''}} is a multiple of {{pi}} and therefore {{math|sin(''nx<sub>r</sub>'') {{=}} 0}}. So, <math display=block>0 = {{2m + 1} \choose 1} \cot^{2m} x_r - {{2m + 1} \choose 3} \cot^{2m - 2} x_r \pm \cdots + (-1)^m{{2m + 1} \choose {2m + 1}}</math> for every {{math|''r'' {{=}} 1, 2, ..., ''m''}}. The values {{math|''x<sub>r</sub>'' {{=}} ''x''<sub>1</sub>, ''x''<sub>2</sub>, ..., ''x<sub>m</sub>''}} are distinct numbers in the interval {{math|0 < {{math|''x<sub>r</sub>''}} < {{sfrac|{{pi}}|2}}}}. Since the function {{math|cot<sup>2</sup> ''x''}} is [[Injective function|one-to-one]] on this interval, the numbers {{math|''t<sub>r</sub>'' {{=}} cot<sup>2</sup> ''x<sub>r</sub>''}} are distinct for {{math|''r'' {{=}} 1, 2, ..., ''m''}}. By the above equation, these {{math|''m''}} numbers are the roots of the {{math|''m''}}th degree polynomial <math display=block>p(t) = {{2m + 1} \choose 1}t^m - {{2m + 1} \choose 3}t^{m - 1} \pm \cdots + (-1)^m{{2m+1} \choose {2m + 1}}.</math> By [[Vieta's formulas]] we can calculate the sum of the roots directly by examining the first two coefficients of the polynomial, and this comparison shows that <math display=block>\cot ^2 x_1 + \cot ^2 x_2 + \cdots + \cot ^2 x_m = \frac{\binom{2m + 1}3} {\binom{2m + 1}1} = \frac{2m(2m - 1)}6.</math> Substituting the [[list of trigonometric identities|identity]] {{math|csc<sup>2</sup> ''x'' {{=}} cot<sup>2</sup> ''x'' + 1}}, we have <math display=block>\csc ^2 x_1 + \csc ^2 x_2 + \cdots + \csc ^2 x_m = \frac{2m(2m - 1)}6 + m = \frac{2m(2m + 2)}6.</math> Now consider the inequality {{math|cot<sup>2</sup> ''x'' < {{sfrac|1|''x''<sup>2</sup>}} < csc<sup>2</sup> ''x''}} (illustrated geometrically above). If we add up all these inequalities for each of the numbers {{math|''x<sub>r</sub>'' {{=}} {{sfrac|''r''{{pi}}|2''m'' + 1}}}}, and if we use the two identities above, we get <math display=block>\frac{2m(2m - 1)}6 < \left(\frac{2m + 1}{\pi} \right)^2 + \left(\frac{2m + 1}{2\pi} \right)^2 + \cdots + \left(\frac{2m + 1}{m \pi} \right)^2 < \frac{2m(2m + 2)}6.</math> Multiplying through by {{math|<big><big>(</big></big>{{sfrac|{{pi}}|2''m'' + 1}}<big><big>)</big></big>{{su|p=2}}}}, this becomes <math display=block>\frac{\pi ^2}{6}\left(\frac{2m}{2m + 1}\right)\left(\frac{2m - 1}{2m + 1}\right) < \frac{1}{1^2} + \frac{1}{2^2} + \cdots + \frac{1}{m^2} < \frac{\pi ^2}{6}\left(\frac{2m}{2m + 1}\right)\left(\frac{2m + 2}{2m + 1}\right).</math> As {{math|''m''}} approaches infinity, the left and right hand expressions each approach {{sfrac|{{pi}}<sup>2</sup>|6}}, so by the [[squeeze theorem]], <math display=block>\zeta(2) = \sum_{k=1}^\infty \frac{1}{k^2} = \lim_{m \to \infty}\left(\frac{1}{1^2} + \frac{1}{2^2} + \cdots + \frac{1}{m^2}\right) = \frac{\pi ^2}{6}</math> and this completes the proof. ==Proof assuming Weil's conjecture on Tamagawa numbers== A proof is also possible assuming [[Weil's conjecture on Tamagawa numbers]].<ref>{{citation|title=Algebraic groups and number theory|translator=Rachel Rowen|publisher=Academic Press|author1=[[Vladimir Platonov]]|author2=Andrei Rapinchuk|year=1994}}|</ref> The conjecture asserts for the case of the [[algebraic group]] SL<sub>2</sub>('''R''') that the [[Tamagawa number]] of the group is one. That is, the quotient of the special linear group over the rational [[Adele ring|adeles]] by the special linear group of the rationals (a [[compact set]], because <math>SL_2(\mathbb Q)</math> is a lattice in the adeles) has Tamagawa measure 1: <math display="block">\tau(SL_2(\mathbb Q)\setminus SL_2(A_{\mathbb Q}))=1.</math> To determine a Tamagawa measure, the group <math>SL_2</math> consists of matrices <math display="block">\begin{bmatrix}x&y\\z&t\end{bmatrix}</math> with <math>xt-yz=1</math>. An invariant [[volume form]] on the group is <math display="block">\omega = \frac1x dx\wedge dy\wedge dz.</math> The measure of the quotient is the product of the measures of <math>SL_2(\mathbb Z)\setminus SL_2(\mathbb R)</math> corresponding to the infinite place, and the measures of <math>SL_2(\mathbb Z_p)</math> in each finite place, where <math>\mathbb Z_p</math> is the [[p-adic integers]]. For the local factors, <math display="block">\omega(SL_2(\mathbb Z_p)) = |SL_2(F_p)|\omega(SL_2(\mathbb Z_p,p))</math> where <math>F_p</math> is the field with <math>p</math> elements, and <math>SL_2(\mathbb Z_p,p)</math> is the [[congruence subgroup]] modulo <math>p</math>. Since each of the coordinates <math>x,y,z</math> map the latter group onto <math>p\mathbb Z_p</math> and <math>\left|\frac1x\right|_p=1</math>, the measure of <math>SL_2(\mathbb Z_p,p)</math> is <math>\mu_p(p\mathbb Z_p)^3=p^{-3}</math>, where <math>\mu_p</math> is the normalized [[Haar measure]] on <math>\mathbb Z_p</math>. Also, a standard computation shows that <math>|SL_2(F_p)|=p(p^2-1)</math>. Putting these together gives <math>\omega(SL_2(\mathbb Z_p))=(1-1/p^2)</math>. At the infinite place, an integral computation over the fundamental domain of <math>SL_2(\mathbb Z)</math> shows that <math>\omega(SL_2(\mathbb Z)\setminus SL_2(\mathbb R)=\pi^2/6</math>, and therefore the Weil conjecture finally gives <math display="block">1 = \frac{\pi^2}6\prod_p \left(1-\frac1{p^2}\right).</math> On the right-hand side, we recognize the [[Euler product]] for <math>1/\zeta(2)</math>, and so this gives the solution to the Basel problem. This approach shows the connection between (hyperbolic) geometry and arithmetic, and can be inverted to give a proof of the Weil conjecture for the special case of <math>SL_2</math>, contingent on an independent proof that <math>\zeta(2)=\pi^2/6</math>. ==Geometric proof== The Basel problem can be proved with [[Euclidean geometry]], using the insight that ''the real line can be seen as a circle of infinite radius''. An intuitive, if not completely rigorous, sketch is given here. * Choose an integer <math>N</math>, and take <math>N</math> equally spaced points on a circle with ''circumference'' equal to <math>2N</math>. The radius of the circle is <math>N/\pi</math> and the length of each [[Circular arc|arc]] between two points is <math>2</math>. Call the points <math>P_{1..N}</math>. * Take another generic point <math>Q</math> on the circle, which will lie at a fraction <math>0 < \alpha < 1</math> of the arc between two consecutive points (say <math>P_1</math> and <math>P_2</math> without loss of generality). * Draw all the [[Chord (geometry)|chords]] joining <math>Q</math> with each of the <math>P_{1..N}</math> points. Now (this is the key to the proof), compute the ''sum of the inverse squares'' of the lengths of all these chords, call it <math>sisc</math>. * The proof relies on the notable fact that (for a fixed <math>\alpha</math>), ''the <math>sisc</math> does not depend on <math>N</math>.'' Note that intuitively, as <math>N</math> increases, the number of chords increases, but their length increases too (as the circle gets bigger), so their inverse square decreases. * In particular, take the case where <math>\alpha = 1/2</math>, meaning that <math>Q</math> is the midpoint of the arc between two consecutive <math>P</math>'s. The <math>sisc</math> can then be found trivially from the case <math>N=1</math>, where there is only one <math>P</math>, and one <math>Q</math> on the opposite side of the circle. Then the chord is the diameter of the circle, of length <math>2/\pi</math>. The <math>sisc</math> is then <math>\pi^2/4</math>. * When <math>N</math> goes to infinity, the circle approaches the real line. If you set the origin at <math>Q</math>, the points <math>P_{1..N}</math> are positioned at the ''odd'' integer positions (positive and negative), since the arcs have length 1 from <math>Q</math> to <math>P_1</math>, and 2 onward. You hence get this variation of the Basel Problem: <math display="block"> \sum_{z=-\infty}^{\infty} \frac{1}{(2z-1)^2} = \frac{\pi^2}{4} </math> * From here, you can recover the original formulation with a bit of algebra, as: <math display="block"> \sum_{n=1}^\infty \frac{1}{n^2} = \sum_{n=1}^\infty \frac{1}{(2n-1)^2} + \sum_{n=1}^\infty \frac{1}{(2n)^2} = \frac{1}{2}\sum_{z=-\infty}^{\infty} \frac{1}{(2z-1)^2} + \frac{1}{4}\sum_{n=1}^\infty \frac{1}{n^2} </math> that is, <math display="block"> \frac{3}{4}\sum_{n=1}^\infty \frac{1}{n^2} = \frac{\pi^2}{8} </math> or <math display="block"> \sum_{n=1}^\infty \frac{1}{n^2} = \frac{\pi^2}{6} </math>. The independence of the <math>sisc</math> from <math>N</math> can be proved easily with Euclidean geometry for the more restrictive case where <math>N</math> is a power of 2, i.e. <math>N = 2^n</math>, which still allows the limiting argument to be applied. The proof proceeds by [[Mathematical induction|induction]] on <math>n</math>, and uses the [[Inverse Pythagorean theorem|Inverse Pythagorean Theorem]], which states that: <math display="block"> \frac{1}{a^2} + \frac{1}{b^2} = \frac{1}{h^2} </math> where <math>a</math> and <math>b</math> are the legs and <math>h</math> is the height of a right triangle. * In the base case of <math>n=0</math>, there is only 1 chord. In the case of <math>\alpha = 1/2</math>, it corresponds to the diameter and the <math>sisc</math> is <math>\pi^2/4</math> as stated above. * Now, assume that you have <math>2^n</math> points on a circle with radius <math>2^n/\pi</math> and center <math>O</math>, and <math>2^{n+1}</math> points on a circle with radius <math>2^{n+1}/\pi</math> and center <math>R</math>. The induction step consists in showing that these 2 circles have the same <math>sisc</math> for a given <math>\alpha</math>. * Start by drawing the circles so that they share point <math>Q</math>. Note that <math>R</math> lies on the smaller circle. Then, note that <math>2^{n+1}</math> is always even, and a simple geometric argument shows that you can pick ''pairs'' of opposite points <math>P_1</math> and <math>P_2</math> on the larger circle by joining each pair with a diameter. Furthermore, for each pair, one of the points will be in the "lower" half of the circle (closer to <math>Q</math>) and the other in the "upper" half. [[File:Induction step 1728668638527.jpg|thumb|The sum of inverse squares of distances of P1 and P2 from Q equals the inverse square distance from P to Q.]] * The diameter of the bigger circle <math>P_1P_2</math> cuts the smaller circle at <math>R</math> and at another point <math>P</math>. You can then make the following considerations: ** <math>P_1 \widehat{Q} P_2</math> is a right angle, since <math>P_1P_2</math> is a diameter. ** <math>Q \widehat{P} R</math> is a right angle, since <math>QR</math> is a diameter. ** <math>Q \widehat{R} P_2 = Q \widehat{R} P</math> is half of <math>Q \widehat{O} P</math> for the [[Inscribed angle|Inscribed Angle Theorem]]. ** Hence, the arc <math>QP</math> is equal to the arc <math>QP_2</math>, again because the radius is half. ** The chord <math>QP</math> is the height of the right triangle <math>QP_1P_2</math>, hence for the Inverse Pythagorean Theorem: <math display="block"> \frac{1}{\overline{QP}^2} = \frac{1}{\overline{QP_1}^2} + \frac{1}{\overline{QP_2}^2} </math> * Hence for half of the points on the bigger circle (the ones in the lower half) there is a corresponding point on the smaller circle with the same arc distance from <math>Q</math> (since the circumference of the smaller circle is half that of the bigger circle, the last two points closer to <math>R</math> must have arc distance 2 as well). Vice versa, for each of the <math>2^n</math> points on the smaller circle, we can build a pair of points on the bigger circle, and all of these points are equidistant and have the same arc distance from <math>Q</math>. * Furthermore, the total <math>sisc</math> for the bigger circle is the same as the <math>sisc</math> for the smaller circle, since each pair of points on the bigger circle has the same inverse square sum as the corresponding point on the smaller circle.<ref>{{cite web |url= https://www.math.chalmers.se/~wastlund/Cosmic.pdf |title= Summing Inverse Squares by Euclidean Geometry |author= Johan Wästlund |date= December 8, 2010 |access-date= 2024-10-11 |website= Chalmers University of Technology |publisher= Department of Mathematics, Chalmers University}}</ref> ==Other identities== See the special cases of the identities for the [[Riemann zeta function#Representations|Riemann zeta function]] when <math>s = 2.</math> Other notably special identities and representations of this constant appear in the sections below. ===Series representations=== The following are series representations of the constant:<ref name="MWZETA2">{{mathworld|title=Riemann Zeta Function \zeta(2)|id=RiemannZetaFunctionZeta2|mode=cs2}}</ref> <math display=block>\begin{align} \zeta(2) &= 3 \sum_{k=1}^\infty \frac{1}{k^2 \binom{2k}{k}} \\[6pt] &= \sum_{i=1}^\infty \sum_{j=1}^\infty \frac{(i-1)! (j-1)!}{(i+j)!}. \end{align}</math> There are also [[Bailey–Borwein–Plouffe formula|BBP-type]] series expansions for {{math|''ζ''(2)}}.<ref name="MWZETA2" /> ===Integral representations=== The following are integral representations of <math>\zeta(2)\text{:}</math><ref>{{cite arXiv|last1=Connon|first1=D. F.|title=Some series and integrals involving the Riemann zeta function, binomial coefficients and the harmonic numbers (Volume I)|year=2007|class=math.HO|eprint=0710.4022|mode=cs2}}</ref><ref>{{mathworld|title=Double Integral|id=DoubleIntegral|mode=cs2}}</ref><ref>{{mathworld|title=Hadjicostas's Formula|id=HadjicostassFormula.html|mode=cs2}}</ref> <math display=block> \begin{align} \zeta(2) & = -\int_0^1 \frac{\log x}{1-x} \, dx \\[6pt] & = \int_0^{\infty} \frac{x}{e^x-1} \, dx \\[6pt] & = \int_0^1 \frac{(\log x)^2}{(1+x)^2} \, dx \\[6pt] & = 2 + 2\int_1^{\infty} \frac{\lfloor x \rfloor -x}{x^3} \, dx \\[6pt] & = \exp\left(2 \int_2^{\infty} \frac{\pi(x)}{x(x^2-1)} \,dx\right) \\[6pt] & = \int_0^1 \int_0^1 \frac{dx \, dy}{1-xy} \\[6pt] & = \frac{4}{3} \int_0^1 \int_0^1 \frac{dx \, dy}{1-(xy)^2} \\[6pt] & = \int_0^1 \int_0^1 \frac{1-x}{1-xy} \, dx \, dy + \frac{2}{3}. \end{align}</math> ===Continued fractions=== In van der Poorten's classic article chronicling [[Apéry's constant|Apéry's proof of the irrationality of <math>\zeta(3)</math>]],<ref>{{Citation |first=Alfred |last=van der Poorten |author-link=Alfred van der Poorten |title=A proof that Euler missed ... Apéry's proof of the irrationality of {{math|''ζ''(3)}} |journal=[[The Mathematical Intelligencer]] |volume=1 |issue=4 |year=1979 |pages=195–203 |doi=10.1007/BF03028234 |s2cid=121589323 |url=http://www.maths.mq.edu.au/~alf/45.pdf |url-status=dead |archive-url=https://web.archive.org/web/20110706114957/http://www.maths.mq.edu.au/~alf/45.pdf |archive-date=2011-07-06 }}</ref> the author notes as "a red herring" the similarity of a [[simple continued fraction]] for Apery's constant, and the following one for the Basel constant: <math display=block>\frac{\zeta(2)}{5} = \cfrac{1}{\widetilde{v}_1 - \cfrac{1^4}{\widetilde{v}_2-\cfrac{2^4}{\widetilde{v}_3-\cfrac{3^4}{\widetilde{v}_4-\ddots}}}}, </math> where <math>\widetilde{v}_n = 11n^2-11n+3 \mapsto \{3,25,69,135,\ldots\}</math>. Another continued fraction of a similar form is:<ref name="Berndt">{{citation |last1=Berndt |first1=Bruce C. |title=Ramanujan's Notebooks: Part II |date=1989 |publisher=Springer-Verlag |isbn=978-0-387-96794-3 |page=150}}</ref> <math display=block>\frac{\zeta(2)}{2} = \cfrac{1}{v_1 - \cfrac{1^4}{v_2-\cfrac{2^4}{v_3-\cfrac{3^4}{v_4-\ddots}}}}, </math> where <math>v_n = 2n-1 \mapsto \{1,3,5,7,9,\ldots\}</math>.
Edit summary
(Briefly describe your changes)
By publishing changes, you agree to the
Terms of Use
, and you irrevocably agree to release your contribution under the
CC BY-SA 4.0 License
and the
GFDL
. You agree that a hyperlink or URL is sufficient attribution under the Creative Commons license.
Cancel
Editing help
(opens in new window)