← Mathematics
The Gauss-Legendre algorithm


Abstract.   We prove the following approximation of $ \pi $

\[\pi = \frac{2 M (1, \sqrt{2}/2)^2}{1 - \displaystyle{\sum_{n=0}^{\infty} 2^n (a_n^2 - b_n^2)}},\]

which is derived from the arithmetic-geometric mean of $ 1 $ and $ \sqrt{2}/2 $.

Note.   Throughout this article tedious steps are often skipped to illustrate the main picture; in particular, computing integral substitutions and power series expansions.

Table of Contents

  1. History
  2. Elliptic integrals
  3. Main results
  4. References

1. History

  The methods used for approximating $ \pi $ span millennia and severely vary in complexity; we will describe a few of the popular methods. The reference here is [1]. Around 200 BC, Archimedes approximated the circumference $ C $ and radius $ r $ of a circle by inscribing it in a polygon with $ n $ sides. It is not hard to see that as $ n $ approaches infinity, we get $ \pi $ using the circumference formula $ C = 2 \pi r $. Then during the invention of calculus in the 1600s, Newton and others used integrals and power series expansions to calculate $ \pi $. For instance, the identity

\[\arctan(x) = \sum_{n=0}^{\infty} (-1)^{n} \frac{x^{2n+1}}{2n+1}\]

evaluated at $ x = 1 $ gives us

\[\frac{\pi}{4} = 1-\frac{1}{3}+\frac{1}{5}+\frac{1}{7}-\cdots.\]

  Furthermore, in the 1700s, Euler calculated the values of the Riemann zeta function. Famously, for $ x=2 $, we get that

\[\zeta(2) = \sum_{n=1}^{\infty} \frac{1}{n^2} = \frac{\pi^2}{6}.\]

The underlying problem is this family of methods are slow, often taking hundreds of iterations to even yield a couple of digits.

  In contrast, the Gauss-Legendre algorithm has quadratic convergence. Let us say we want to calculate 512 decimal places of $ \pi $. Then the Gauss-Legendre algorithm only needs 9 iterations, while almost all of the older methods (normally) need at least 800 iterations, if not significantly more. As an aside, Ramanujan's equation for $ \pi $

\[\frac{1}{\pi} = \frac{2 \sqrt{2}}{9801} \sum_{n=0}^{\infty} \frac{(4n)! (1103+26390n)}{(n!)^4 396^{4n}}.\]

only need around $ 65 $ iterations to converge [3]. His equations are now used for large approximations of $ \pi $ due to computational complexity optimization and storage restrictions.

2. Elliptic integrals

  Let $ (a \cos \theta, b \sin \theta) $ be an ellipse parameterized by $ \theta \in [0, 2 \pi] $. Then it's arc-length is given by

\[\int_{0}^{2 \pi} \sqrt{a^2 \cos^2 \theta + b^2 \sin^2 \theta} d \theta. \tag{1}\]

This integral is not easy to compute in itself, and it was generalized to the study of so-called elliptic integrals in the early 1700s. Now notice that in (1) all of the data concerning our specific ellipse's arc-length is contained within the interval $ 0 \leq \theta \leq \pi/2 $. Hence, it would make sense to reduce our study to complete elliptic integrals, meaning those with amplitude $ \pi/2 $. Furthermore, we will restrict ourselves to complete elliptic integrals of the first and second kind. For the proof that every elliptic integral is of the first, second, or third kind, see [2].

Definition.   Let

\[\begin{aligned} F(k) & = \int_0^{\pi/2} \frac{1}{\sqrt{1 - k^2 \sin^2 \theta}} d \theta, \\ E(k) & = \int_0^{\pi/2} \sqrt{1 - k^2 \sin^2 \theta} d \theta, \\ F_S(a, b) & = \int_0^{\pi/2} \frac{1}{\sqrt{a^2 \cos^2 \theta + b^2 \sin^2 \theta}} d \theta, \\ E_S(a, b) & = \int_0^{\pi/2} \sqrt{a^2 \cos^2 \theta + b^2 \sin^2 \theta} d \theta. \end{aligned}\]

Then $ F(k) $ and $ E(k) $ are called complete elliptic integrals of the first and second kinds, respectively, with symmetric forms $ F_S(a, b) $ and $ E_S(a, b) $. We refer to $ k $ as the modulus of our integral.

  From here on we will assume that by elliptic integral we mean a complete elliptic integral of the first or second kind. We see that the ordinary and symmetric forms of elliptic integrals are related by the following proposition.

Proposition 2.1.   Let $ k^2 = 1 - b^2/a^2 $. Then the identities

\[\begin{aligned} F(k) & = a F_S(a, b) \\ E(k) & = \frac{1}{a} E_S(a, b). \end{aligned}\]

are true.

Proof. Let us prove the second equation with the first following similarly. Substituting, we get

\[\begin{aligned} E(k) & = \int_0^{\pi/2} \sqrt{1 - k^2 \sin^2 \theta} d \theta \\ & = \int_0^{\pi/2} \sqrt{1 - (1-b^2/a^2) \sin^2 \theta} d \theta \\ & = \int_0^{\pi/2} \sqrt{\cos^2 \theta + b^2/a^2 \sin^2 \theta} d \theta \\ & = \frac{1}{a} \int_0^{\pi/2} \sqrt{a^2 \cos^2 \theta + b^2 \sin^2 \theta} d \theta \\ & = \frac{1}{a} E_S(a, b), \end{aligned}\]

and hence our result. $ \blacksquare $

  We also recall a result regarding the symmetry of these integrals originally discovered by Legendre.

Proposition 2.2. (Legendre's Identity).   Suppose $ k_1^2 + k_2^2 = 1 $. Then

\[F(k_1) E(k_2) + F(k_2) E(k_1) - F(k_1) F(k_2) = \frac{\pi}{2}\]

holds.

Proof. We leave the details of this proof to the reader. Taking the derivative with respect to $ k_1 $ shows that the L.H.S. is constant. To see that this value is $ \pi/2 $, we take the limit as $ k_1 $ goes to $ 0 $. $ \blacksquare $

3. Main results

  We now prove the Gauss-Legendre algorithm. We will not discuss error analysis, which is done in [4].

Definition.   Let $ a_0, b_0 \in \mathbb{R} $. Let $ a_{n+1} = (a_n + b_n)/2 $ and $ b_{n+1} = (a_n b_n)^{1/2} $ be the arithmetic and geometric means, respectively, of the $ n $th terms. Then we call their common limit

\[\lim_{n \to \infty} a_n = \lim_{n \to \infty} b_n = M(a_0, b_0)\]

the arithmetic-geometric (AM-GM) mean of $ a_0 $ and $ b_0 $.

Theorem (Gauss-Legendre).   Set $ a_0 = 1 $ and $ b_0 = \sqrt{2}/2 $. Then the series

\[\pi = \frac{2 M (1, \sqrt{2}/2)^2}{1 - \displaystyle{\sum_{n=0}^{\infty} 2^n (a_n^2 - b_n^2)}}\]

converges to $ \pi $.

  We need two lemmas in order to prove our result.

Lemma 3.1.   Let $ a_0 = a $ and $ b_0 = b $ as in definition 3.1. Set

\[S = a^2 - \sum_{n=0}^{\infty} 2^{n-1} \left( a_n^2 - b_n^2 \right).\]

Then $ E_S(a, b) = S F_S(a, b) $.

Proof. This proof is taken from [5]. Consider the integral

\[L(a, b) = a^2 F_S(a, b) - E_S(a, b) \tag{2}\]

Explicitly, this expands to

\[L(a, b) = (a^2 - b^2) \int_0^{\pi/2} \frac{\sin^2 \theta}{\sqrt{a^2 \cos^2 \theta + b^2 \sin^2 \theta}} d \theta.\]

  Substituting $ x^2 = a^2 \cos^2 \theta + b^2 \sin^2 \theta $,

\[L(a, b) = \int_b^a \sqrt{\frac{a^2-x^2}{x^2-b^2}} dx.\]

Now substituting $ y = (x+ab/x)/2 $ and considering the associated AM-GM sequence,

\[\begin{aligned} L(a, b) & = \frac{1}{2} \int_{b_1}^{a_1} \frac{(a^2 - b^2) + 4(a_1^2 - y^2)}{\sqrt{(a_1^2 - y^2)(y^2 - b_1^2)}} dy \\ & = \frac{1}{2} (a^2 - b^2) F_S(a, b) + 2 L(a_1, b_1), \end{aligned}\]

and thus

\[\frac{L(a, b)}{F_S(a, b)} = \frac{1}{2}(a_0^2 - b_0^2) + 2 \frac{L(a_1, b_1)}{F_S(a, b)}.\]

  Since $ 2^n(a_n^2 - b_n^2) \to 0 $, we get $ 2^n L(a_n, b_n) \to 0 $, and hence repeatedly applying this identity we get

\[L(a, b) = \sum_{n=0}^{\infty} 2^{n-1} (a_n^2 - b_n^2) F_S(a, b),\]

or equivalently

\[L(a, b) = - S F_S(a, b) + a^2 F_S(a, b).\]

Combining this equation with (2) gives us the result. $ \blacksquare $

Lemma 3.2.   We have

\[F_S(a, b) = \frac{\pi}{2 M(a, b)}\]

In particular,

\[F_S(1, \sqrt{2}) = \frac{\pi}{2 M(1, \sqrt{2})}.\]

Proof. It is not hard to calculate the power series expansion

\[F(k) = \frac{\pi}{2M(1+k, 1-k)} = \sum_{n=0}^{\infty} \left( \frac{(2n-1)!!}{(2n)!!} k^n \right)^2.\]

(The "!!" means double factorial.) Calculating the AM and GM means gives us

\[M(1+k, 1-k) = M(1, \sqrt{1-k^2}).\]

Hence, applying proposition 2.1 for $ b/a = \sqrt{1-k^2} $, we get

\[a F_S(a, b) = \frac{\pi}{2 M(1, b/a)} \\ F_S(a, b) = \frac{\pi}{2 M(a, b)},\]

with the last equality following from the identity $ M(ca, cb) = cM(a, b) $. $ \blacksquare $

  We finally have the tools we need to prove the Gauss-Legendre Algorithm does indeed converge to $ \pi $.

Proof (Gauss-Legendre). Set $ k = \sqrt{2}/2 $ to be our modulus. Then we notice $ 2k^2 = 1 $, and hence we can apply proposition 2.2 to get

\[2 F (k) E (k) - F(k)^2 = \frac{\pi}{2}.\]

Now let us evaluate these integrals by first converting them into symmetric form then applying our lemmas. Since $ k^2 = 1 - 1/k^2 $, we can set $ a = 1 $ and $ b = k $ in proposition 2.1 to get

\[2 F_S(1, k) E_S(1, k) - F_S(1, k)^2 = \frac{\pi}{2}.\]

Then applying lemma 3.1 gives us an equation only dependent on $ F_S(1, k) $

\[(2 S - 1) F_S(1, k)^2 = \frac{\pi}{2}.\]

Finally, we apply lemma 3.2 to write our equation in terms of the AM-GM mean

\[\begin{aligned} \frac{\pi}{2} & = (2S-1) \left( \frac{\pi}{2 M(1, k)} \right)^2 \\ \pi & = \frac{2M(1, k)^2}{2S-1}. \end{aligned}\]

Plugging in $ S $ gives us our result. $ \blacksquare $

4. References

  1. David H. Bailey, Simon M. Plouffe, Peter B. Borwein, and Jonathan M. Borwein, The quest for $ \pi $, The Mathematical Intelligencer 19 (1997), no. 1, 50–56.
  2. Gosta Mittag-Leffler, An introduction to the theory of elliptic functions, Annals of Mathematics 24 (1923), no. 4, 271–351.
  3. Srinivasa Ramanujan, Modular equations and approximations to $ \pi $, The Quarterly Joural of Pure and Applied Mathematics 45 (1914), 350372.
  4. Eugene Salamin, Computation of $ \pi $ using arithmetic-geometric mean, Mathematics of Computation 30 (1976), no. 135, 565–570.
  5. Paramanand Singh, $ \pi $ and the AGM: Evaluating elliptic integrals, 2009. Link.