11

I'm interested in the following definite integral: $$\int_0^\infty\big(2J_0(2x)^2-J_0(x)^2+2Y_0(2x)^2-Y_0(x)^2\big)\,dx,\tag1$$ where $J_\nu$ and $Y_\nu$ are the Bessel functions of the first and the second kind.

Mathematica evaluates this integral symbolically to $\frac{\ln2}\pi$, but the result of a numerical integration looks more like $\frac{\ln4}\pi$, so I suspect the symbolic result is incorrect.

Moreover, it looks like both components converge and make equal contribution: $$\int_0^\infty\big(2J_0(2x)^2-J_0(x)^2\big)\,dx\stackrel?=\int_0^\infty\big(2Y_0(2x)^2-Y_0(x)^2\big)\,dx\stackrel?=\frac{\ln2}\pi,\tag2$$ but it is more difficult to check numerically, because both integrands here are oscillating (unlike $(1)$ where the integrand looks monotonic).

How can we find values of these integrals and prove them correct? Can we generalize results for values of the index $\nu$ other than $0$?

  • 1
    When a symbolic integrator returns an incorrect value for a high-brow integral of special functions that just happens to be off by an exact integer factor, I would suspect a discrepancy in the internal function definitions being called. – David H May 30 '16 at 02:52
  • 1
    It is useful to exploit the fact that the Laplace transform of $J_0^2$ is given by a complete elliptic integral of the first kind. I am able to prove that $\int_{0}^{+\infty}\left( 2J_0(2x)^2-J_0(x)^2\right),dx = \frac{\log 2}{\pi}$ this way, but the $Y_0$ function still eludes me. I will work on this tomorrow. – Jack D'Aurizio May 30 '16 at 03:13
  • I suspect that the same result holds for $0\le\nu<1/2$. – Vladimir Reshetnikov May 30 '16 at 03:48

2 Answers2

15

I stated an interesting thing in the comments (wisely explored by Random Variable), but it is definitely not the fastest way to go.

For the glory of Italian mathematicians, Frullani's theorem is the way (again):

$$ \int_{0}^{+\infty}\frac{x\, J_0(x)^2-2x\, J_0(2x)^2}{x}\,dx = (-\log 2)\cdot\underset{x\to+\infty}{\widetilde{\lim}} x\cdot J_0(x)^2 = \color{red}{-\frac{\log 2}{\pi}} $$ and the argument is just the same for the $Y_0$ function.

The wide tilde-limit has to be intended as a Cesàro mean: $$\underset{x\to+\infty}{\widetilde{\lim}}f(x) = \lim_{x\to +\infty}\frac{1}{x}\int_{x}^{2x}f(z)\,dz.$$

Moreover, provided that the integral converges, the previous result holds unchanged also by replacing $J_0$ with $J_\nu$ for any $\nu\geq 0$, since: $$ J_\nu(z) = \sqrt{\frac{2}{\pi z}}\left(\cos \left(z-\frac{\nu\pi}{2}-\frac{\pi}{4}\right)+O\left(\frac{1}{z}\right)\right) $$ for any large $z\in\mathbb{R}^+$.

Jack D'Aurizio
  • 353,855
  • Wait, why is Frullani's theorem applicable to a Cesàro mean limit? – Vladimir Reshetnikov May 30 '16 at 04:26
  • @VladimirReshetnikov: because Frullani's theorem is just a consequence of integration by parts. To say the truth, I am using a slightly strengthened version of the linked result, that allows us to compute also integrals like $$\int_{0}^{+\infty}\frac{\cos^2(ax)-\cos^2(bx)}{x},dx $$ where the function of interest is oscillating. – Jack D'Aurizio May 30 '16 at 04:32
  • Anyhow, differentiation under the integral sign / Feynman's trick also proves that the main equality in my answer holds true. – Jack D'Aurizio May 30 '16 at 04:34
  • There is also an asymptotic expansion argument: for moderately large values of $x$, $x\cdot J_0(x)^2$ and $\frac{1+\sin(2x)}{\pi}$ are almost the same function. – Jack D'Aurizio May 30 '16 at 04:37
  • @JackD'Aurizio You wrote: "the … result holds unchanged … for any $\nu\ge0$". It looks that integrals with $Y_\nu$ terms for $\nu\ge1/2$ become divergent near the lower bound $x=0$. You probably meant to add "provided that the integral converges". This is a prerequisite for applying Frullani's theorem, right? – Vladimir Reshetnikov May 30 '16 at 15:16
  • @VladimirReshetnikov: oh, sure, you are completely right. – Jack D'Aurizio May 30 '16 at 16:16
  • @JackD'Aurizio The Cesàro mean come from the form of the Frullani's theorem you used, am I right? – Marco Cantarini May 31 '16 at 07:37
  • @MarcoCantarini: yes, you are. – Jack D'Aurizio May 31 '16 at 10:33
8

UPDATE: I added a proof showing that $$\int_{0}^{\infty} \left(2J_{0}(2x)^{2} -J_{0}(x)^2 \right) \, \mathrm dx= \int_{0}^{\infty} \left(2Y_{0}(2x)^2 - Y_{0}(x)^{2} \right) \, \mathrm dx. $$


We can use Ramanujan's master theorem to show that $$\int_{0}^{\infty}\left(2J_0(2x)^2-J_0(x)^2\right)\, \mathrm dx = \frac{\ln 2}{\pi}. $$

As $x \to \infty$, $$\left(2J_0(2x)^2-J_0(x)^2\right) \sim \frac{\sin (4x)-\sin (2x)}{\pi x}. $$ So the integral converges conditionally but not absolutely.

The hypergeometric representation of the square of the Bessel function of the first kind of order zero is $$J_{0}(z)^{2} = \, _1F_2\left(\frac{1}{2}; 1, 1; -z^{2} \right). $$

So for $a>0$ and $0 < s < 1$, the Mellin transform of $J_{0}(ax)^{2}$ is $$ \begin{align} \int_{0}^{\infty} x^{s-1} J_{0}(ax)^{2} \, \mathrm dx &= \int_{0}^{\infty} x^{s-1} \, _1F_2\left(\frac{1}{2}; 1, 1; -(ax)^{2} \right) \, \mathrm dx \\ &= \frac{1}{2a^{s}}\int_{0}^{\infty} u^{s/2-1} \, _1F_2\left(\frac{1}{2}; 1, 1; -u \right) \, \mathrm du \\ &= \frac{1}{2a^{s}} \Gamma\left(\frac{s}{2} \right) \frac{\Gamma \left(\frac{1}{2}-\frac{s}{2} \right) \Gamma(1)^{2} }{\Gamma\left(\frac{1}{2}\right) \Gamma\left(1- \frac{s}{2} \right)^{2}} \\ &= \frac{1}{2a^{s}} \Gamma\left(\frac{s}{2} \right) \frac{\Gamma \left(\frac{1}{2}-\frac{s}{2} \right)}{\sqrt{\pi} \, \Gamma\left(1- \frac{s}{2} \right)^{2}} . \end{align}$$

Since $ \int_{0}^{\infty}\left(2J_0(2x)^2-J_0(x)^2\right)\, \mathrm dx $ converges, a property of the Mellin transform allows us to conclude that $$ \begin{align} \lim_{s \to 1^{-}} \int_{0}^{\infty} x^{s-1} \left(2J_0(2x)^2-J_0(x)^2\right)\, \mathrm dx &= \int_{0}^{\infty} \left(2J_0(2x)^2-J_0(x)^2\right)\, \mathrm dx \\ &= \frac{1}{\pi} \lim_{s \to 1^{-}} \left(\frac{1}{2^{s}}- \frac{1}{2} \right) \Gamma \left(\frac{1}{2} - \frac{s}{2} \right) \\ &= \frac{1}{\pi} \lim_{s \to 0^{-}} \left( \frac{1}{2^{s+1}} - \frac{1}{2} \right)\Gamma \left(- \frac{s}{2} \right) \\ &= \frac{1}{\pi} \lim_{s \to 0^{-}} \left(- \frac{\ln 2}{2} s + \mathcal{O}(s^{2}) \right) \left(- \frac{2}{s} + \mathcal{O}(1) \right) \\ &= \frac{\ln 2}{\pi}. \end{align}$$


Let $H_{0}^{(1)}(z) $ be the Hankel function of the first kind of order zero defined as $$H_{0}^{(1)}(z) = J_{0}(z)+iY_{0}(z). $$

To show that $$\int_{0}^{\infty}\left(2J_0(2x)^2-J_0(x)^2\right)\, \mathrm dx$$ and $$\int_{0}^{\infty}\left(2Y_0(2x)^2-Y_0(x)^2\right)\, \mathrm dx $$ have the same value, we can integrate the function $$2\left( H_{0}^{(1)}(2z)\right)^{2}-\left(H_{0}^{(1)}(z) \right)^{2}$$ around a closed, semicircular contour in the upper half-plane that is indented at the origin.

Applying Jordan's lemma, we get $$\int_{0}^{\infty} \left( 2\left( H_{0}^{(1)}(2xe^{i \pi})\right)^{2}-\left(H_{0}^{(1)}(xe^{i \pi} ) \right)^{2} + 2\left( H_{0}^{(1)}(2x)\right)^{2}-\left(H_{0}^{(1)}(x) \right)^{2} \right) \, \mathrm dx = 0. \tag{1}$$

Since $$H_{0}^{(1)}(xe^{i \pi}) = -J_{0}(x)+iY_{0}(x) \, , \quad x >0, $$ we can rewrite $(1)$ as $$2 \int_{0}^{\infty} \left(2J_{0}(2x)^{2} -J_{0}(x)^2-2Y_{0}(2x)^2 + Y_{0}(x)^{2} \right) = 0, $$ and the result follows.

  • (+1) Very interesting approach. What about the $Y_0^2$ function? – Jack D'Aurizio May 30 '16 at 18:01
  • @JackD'Aurizio Thanks. I'm still trying to figure out how to deal with the other integral. – Random Variable May 30 '16 at 18:28
  • When persuing almost the same path, I got that the Laplace transform of $Y_0^2$ is given by the difference between an elliptic integral and a $$\phantom{}_3 F_2\left(1,1,1;\frac{3}{2},\frac{3}{2};z\right)$$ function. Maybe that helps and sheds some light on peculiar properties of such a hypergeometric function. – Jack D'Aurizio May 30 '16 at 18:30