Bussgang theorem

From HandWiki

In mathematics, the Bussgang theorem is a theorem of stochastic analysis. The theorem states that the cross-correlation between a Gaussian signal before and after it has passed through a nonlinear operation are equal to the signals auto-correlation up to a constant. It was first published by Julian J. Bussgang in 1952 while he was at the Massachusetts Institute of Technology.[1]

Statement

Let [math]\displaystyle{ \left\{X(t)\right\} }[/math] be a zero-mean stationary Gaussian random process and [math]\displaystyle{ \left \{ Y(t) \right\} = g(X(t)) }[/math] where [math]\displaystyle{ g(\cdot) }[/math] is a nonlinear amplitude distortion.

If [math]\displaystyle{ R_X(\tau) }[/math] is the autocorrelation function of [math]\displaystyle{ \left\{ X(t) \right\} }[/math], then the cross-correlation function of [math]\displaystyle{ \left\{ X(t) \right\} }[/math] and [math]\displaystyle{ \left\{ Y(t) \right\} }[/math] is

[math]\displaystyle{ R_{XY}(\tau) = CR_X(\tau), }[/math]

where [math]\displaystyle{ C }[/math] is a constant that depends only on [math]\displaystyle{ g(\cdot) }[/math].

It can be further shown that

[math]\displaystyle{ C = \frac{1}{\sigma^3\sqrt{2\pi}}\int_{-\infty}^\infty ug(u)e^{-\frac{u^2}{2\sigma^2}} \, du. }[/math]

Derivation for One-bit Quantization

It is a property of the two-dimensional normal distribution that the joint density of [math]\displaystyle{ y_1 }[/math] and [math]\displaystyle{ y_2 }[/math] depends only on their covariance and is given explicitly by the expression

[math]\displaystyle{ p(y_1,y_2) = \frac{1}{2 \pi \sqrt{1-\rho^2}} e^{-\frac{y_1^2 + y_2^2 - 2 \rho y_1 y_2}{2(1-\rho^2)}} }[/math]

where [math]\displaystyle{ y_1 }[/math] and [math]\displaystyle{ y_2 }[/math] are standard Gaussian random variables with correlation [math]\displaystyle{ \phi_{y_1y_2}=\rho }[/math].

Assume that [math]\displaystyle{ r_2 = Q(y_2) }[/math], the correlation between [math]\displaystyle{ y_1 }[/math] and [math]\displaystyle{ r_2 }[/math] is,

[math]\displaystyle{ \phi_{y_1r_2} = \frac{1}{2 \pi \sqrt{1-\rho^2}} \int_{-\infty}^{\infty} \int_{-\infty}^{\infty} y_1 Q(y_2) e^{-\frac{y_1^2 + y_2^2 - 2 \rho y_1 y_2}{2(1-\rho^2)}} \, dy_1 dy_2 }[/math].

Since

[math]\displaystyle{ \int_{-\infty}^{\infty} y_1 e^{-\frac{1}{2(1-\rho^2)} y_1^2 + \frac{\rho y_2}{1-\rho^2} y_1 } \, dy_1 = \rho \sqrt{2 \pi (1-\rho^2)} y_2 e^{ \frac{\rho^2 y_2^2}{2(1-\rho^2)} } }[/math],

the correlation [math]\displaystyle{ \phi_{y_1 r_2} }[/math] may be simplified as

[math]\displaystyle{ \phi_{y_1 r_2} = \frac{\rho}{\sqrt{2 \pi}} \int_{-\infty}^{\infty} y_2 Q(y_2) e^{-\frac{y_2^2}{2}} \, dy_2 }[/math].

The integral above is seen to depend only on the distortion characteristic [math]\displaystyle{ Q() }[/math] and is independent of [math]\displaystyle{ \rho }[/math].

Remembering that [math]\displaystyle{ \rho=\phi_{y_1 y_2} }[/math], we observe that for a given distortion characteristic [math]\displaystyle{ Q() }[/math], the ratio [math]\displaystyle{ \frac{\phi_{y_1 r_2}}{\phi_{y_1 y_2}} }[/math] is [math]\displaystyle{ K_Q=\frac{1}{\sqrt{2\pi}} \int_{-\infty}^{\infty} y_2 Q(y_2) e^{-\frac{y_2^2}{2}} \, dy_2 }[/math].

Therefore, the correlation can be rewritten in the form

[math]\displaystyle{ \phi_{y_1 r_2} = K_Q \phi_{y_1 y_2} }[/math].

The above equation is the mathematical expression of the stated "Bussgang‘s theorem".

If [math]\displaystyle{ Q(x) = \text{sign}(x) }[/math], or called one-bit quantization, then [math]\displaystyle{ K_Q= \frac{2}{\sqrt{2\pi}} \int_{0}^{\infty} y_2 e^{-\frac{y_2^2}{2}} \, dy_2 = \sqrt{\frac{2}{\pi}} }[/math].

[2][3][1][4]

Arcsine law

If the two random variables are both distorted, i.e., [math]\displaystyle{ r_1 = Q(y_1), r_2 = Q(y_2) }[/math], the correlation of [math]\displaystyle{ r_1 }[/math] and [math]\displaystyle{ r_2 }[/math] is

[math]\displaystyle{ \phi_{r_1 r_2}=\int_{-\infty}^{\infty}\int_{-\infty}^{\infty} Q(y_1) Q(y_2) p(y_1, y_2) \, dy_1 dy_2 }[/math].

When [math]\displaystyle{ Q(x) = \text{sign}(x) }[/math], the expression becomes,

[math]\displaystyle{ \phi_{r_1 r_2}=\frac{1}{2\pi \sqrt{1-\rho^2}} \left[ \int_{0}^{\infty} \int_{0}^{\infty} e^{-\alpha} \, dy_1 dy_2 + \int_{-\infty}^{0} \int_{-\infty}^{0} e^{-\alpha} \, dy_1 dy_2 - \int_{0}^{\infty} \int_{-\infty}^{0} e^{-\alpha} \, dy_1 dy_2 - \int_{-\infty}^{0} \int_{0}^{\infty} e^{-\alpha} \, dy_1 dy_2 \right] }[/math]

where [math]\displaystyle{ \alpha = \frac{y_1^2 + y_2^2 - 2\rho y_1 y_2}{2 (1-\rho^2)} }[/math].

Noticing that

[math]\displaystyle{ \int_{-\infty}^{\infty} \int_{-\infty}^{\infty} p(y_1,y_2) \, dy_1 dy_2 = \frac{1}{2\pi \sqrt{1-\rho^2}} \left[ \int_{0}^{\infty} \int_{0}^{\infty} e^{-\alpha} \, dy_1 dy_2 + \int_{-\infty}^{0} \int_{-\infty}^{0} e^{-\alpha} \, dy_1 dy_2 + \int_{0}^{\infty} \int_{-\infty}^{0} e^{-\alpha} \, dy_1 dy_2 + \int_{-\infty}^{0} \int_{0}^{\infty} e^{-\alpha} \, dy_1 dy_2 \right]=1 }[/math],

and [math]\displaystyle{ \int_{0}^{\infty} \int_{0}^{\infty} e^{-\alpha} \, dy_1 dy_2 = \int_{-\infty}^{0} \int_{-\infty}^{0} e^{-\alpha} \, dy_1 dy_2 }[/math], [math]\displaystyle{ \int_{0}^{\infty} \int_{-\infty}^{0} e^{-\alpha} \, dy_1 dy_2 = \int_{-\infty}^{0} \int_{0}^{\infty} e^{-\alpha} \, dy_1 dy_2 }[/math],

we can simplify the expression of [math]\displaystyle{ \phi_{r_1r_2} }[/math] as

[math]\displaystyle{ \phi_{r_1 r_2}=\frac{4}{2\pi \sqrt{1-\rho^2}} \int_{0}^{\infty} \int_{0}^{\infty} e^{-\alpha} \, dy_1 dy_2-1 }[/math]

Also, it is convenient to introduce the polar coordinate [math]\displaystyle{ y_1 = R \cos \theta, y_2 = R \sin \theta }[/math]. It is thus found that

[math]\displaystyle{ \phi_{r_1 r_2} =\frac{4}{2\pi \sqrt{1-\rho^2}} \int_{0}^{\pi/2} \int_{0}^{\infty} e^{-\frac{R^2 - 2R^2 \rho \cos \theta \sin \theta \ }{2(1-\rho^2)}} R \, dR d\theta-1=\frac{4}{2\pi \sqrt{1-\rho^2}} \int_{0}^{\pi/2} \int_{0}^{\infty} e^{-\frac{R^2 (1-\rho \sin 2\theta )}{2(1-\rho^2)}} R \, dR d\theta -1 }[/math].

Integration gives

[math]\displaystyle{ \phi_{r_1 r_2}=\frac{2\sqrt{1-\rho^2}}{\pi} \int_{0}^{\pi/2} \frac{d\theta}{1-\rho \sin 2\theta} - 1= - \frac{2}{\pi} \arctan \left( \frac{\rho-\tan\theta} {\sqrt{1-\rho^2}} \right) \Bigg|_{0}^{\pi/2} -1 =\frac{2}{\pi} \arcsin(\rho) }[/math]

This is called "Arcsine law", which was first found by J. H. Van Vleck in 1943 and republished in 1966.[2][3] The "Arcsine law" can also be proved in a simpler way by applying Price's Theorem.[4][5]

The function [math]\displaystyle{ f(x)=\frac{2}{\pi} \arcsin x }[/math] can be approximated as [math]\displaystyle{ f(x) \approx \frac{2}{\pi} x }[/math] when [math]\displaystyle{ x }[/math] is small.

Price's Theorem

Given two jointly normal random variables [math]\displaystyle{ y_1 }[/math] and [math]\displaystyle{ y_2 }[/math] with joint probability function

[math]\displaystyle{ {\displaystyle p(y_{1},y_{2})={\frac {1}{2\pi {\sqrt {1-\rho ^{2}}}}}e^{-{\frac {y_{1}^{2}+y_{2}^{2}-2\rho y_{1}y_{2}}{2(1-\rho ^{2})}}}} }[/math],

we form the mean

[math]\displaystyle{ I(\rho)=E(g(y_1,y_2))=\int_{-\infty}^{+\infty} \int_{-\infty}^{+\infty} g(y_1, y_2) p(y_1, y_2) \, dy_1 dy_2 }[/math]

of some function [math]\displaystyle{ g(y_1,y_2) }[/math] of [math]\displaystyle{ (y_1, y_2) }[/math]. If [math]\displaystyle{ g(y_1, y_2) p(y_1, y_2) \rightarrow 0 }[/math] as [math]\displaystyle{ (y_1, y_2) \rightarrow 0 }[/math], then

[math]\displaystyle{ \frac{\partial^n I(\rho)}{\partial \rho^n}=\int_{-\infty}^{\infty} \int_{-\infty}^{\infty} \frac{\partial ^{2n} g(y_1, y_2)}{\partial y_1^n \partial y_2^n} p(y_1, y_2) \, dy_1 dy_2 =E \left(\frac{\partial ^{2n} g(y_1, y_2)}{\partial y_1^n \partial y_2^n} \right) }[/math].

Proof. The joint characteristic function of the random variables [math]\displaystyle{ y_1 }[/math] and [math]\displaystyle{ y_2 }[/math] is by definition the integral

[math]\displaystyle{ \Phi(\omega_1, \omega_2)=\int_{-\infty}^{\infty}\int_{-\infty}^{\infty} p(y_1, y_2) e^{j (\omega_1 y_1 + \omega_2 y_2 )} \, dy_1 dy_2 = \exp \left\{-\frac{\omega_1^2 + \omega_2^2 + 2\rho \omega_1 \omega_2}{2} \right\} }[/math].

From the two-dimensional inversion formula of Fourier transform, it follows that

[math]\displaystyle{ p(y_1, y_2) = \frac{1}{4 \pi^2} \int_{-\infty}^{\infty} \int_{-\infty}^{\infty} \Phi(\omega_1, \omega_2) e^{-j (\omega_1 y_1 + \omega_2 y_2)} \, d\omega_1 d\omega_2 =\frac{1}{4 \pi^2} \int_{-\infty}^{\infty} \int_{-\infty}^{\infty} \exp \left\{-\frac{\omega_1^2 + \omega_2^2 + 2\rho \omega_1 \omega_2}{2} \right\} e^{-j (\omega_1 y_1 + \omega_2 y_2)} \, d\omega_1 d\omega_2 }[/math].

Therefore, plugging the expression of [math]\displaystyle{ p(y_1, y_2) }[/math] into [math]\displaystyle{ I(\rho) }[/math], and differentiating with respect to [math]\displaystyle{ \rho }[/math], we obtain

[math]\displaystyle{ \begin{align} \frac{\partial^n I(\rho)}{\partial \rho^n} & = \int_{-\infty}^{\infty} \int_{-\infty}^{\infty} g(y_1, y_2) p(y_1, y_2) \, dy_1 dy_2 \\ & = \int_{-\infty}^{\infty} \int_{-\infty}^{\infty} g(y_1, y_2) \left(\frac{1}{4 \pi^2} \int_{-\infty}^{\infty} \int_{-\infty}^{\infty} \frac{\partial^ {n}\Phi(\omega_1, \omega_2)}{\partial \rho^n} e^{-j(\omega_1 y_1 + \omega_2 y_2)} \, d\omega_1 d\omega_2 \right) \, dy_1 dy_2 \\ & = \int_{-\infty}^{\infty} \int_{-\infty}^{\infty} g(y_1, y_2) \left(\frac{(-1)^n}{4 \pi^2} \int_{-\infty}^{\infty} \int_{-\infty}^{\infty} \omega_1^n \omega_2^n \Phi(\omega_1, \omega_2) e^{-j(\omega_1 y_1 + \omega_2 y_2)} \, d\omega_1 d\omega_2 \right) \, dy_1 dy_2 \\ & = \int_{-\infty}^{\infty} \int_{-\infty}^{\infty} g(y_1, y_2) \left(\frac{1}{4 \pi^2} \int_{-\infty}^{\infty} \int_{-\infty}^{\infty} \Phi(\omega_1, \omega_2) \frac{\partial^{2n} e^{-j(\omega_1 y_1 + \omega_2 y_2)}}{\partial y_1^n \partial y_2^n} \, d\omega_1 d\omega_2 \right) \, dy_1 dy_2 \\ & = \int_{-\infty}^{\infty} \int_{-\infty}^{\infty} g(y_1, y_2) \frac{\partial^{2n} p(y_1, y_2)}{\partial y_1^n \partial y_2^n} \, dy_1 dy_2 \\ \end{align} }[/math]

After repeated integration by parts and using the condition at [math]\displaystyle{ \infty }[/math], we obtain the Price's theorem.

[math]\displaystyle{ \begin{align} \frac{\partial^n I(\rho)}{\partial \rho^n} & = \int_{-\infty}^{\infty} \int_{-\infty}^{\infty} g(y_1, y_2) \frac{\partial^{2n} p(y_1, y_2)}{\partial y_1^n \partial y_2^n} \, dy_1 dy_2 \\ & = \int_{-\infty}^{\infty} \int_{-\infty}^{\infty} \frac{\partial^{2} g(y_1, y_2)}{\partial y_1 \partial y_2} \frac{\partial^{2n-2} p(y_1, y_2)}{\partial y_1^{n-1} \partial y_2^{n-1}} \, dy_1 dy_2 \\ &=\cdots \\ &=\int_{-\infty}^{\infty} \int_{-\infty}^{\infty} \frac{\partial ^{2n} g(y_1, y_2)}{\partial y_1^n \partial y_2^n} p(y_1, y_2) \, dy_1 dy_2 \end{align} }[/math]

[4][5]

Proof of Arcsine law by Price's Theorem

If [math]\displaystyle{ g(y_1, y_2) = \text{sign}(y_1) \text{sign} (y_2) }[/math], then [math]\displaystyle{ \frac{\partial^2 g(y_1, y_2)}{\partial y_1 \partial y_2} = 4 \delta(y_1) \delta(y_2) }[/math] where [math]\displaystyle{ \delta() }[/math] is the Dirac delta function.

Substituting into Price's Theorem, we obtain,

[math]\displaystyle{ \frac{\partial E(\text{sign} (y_1) \text{sign}(y_2))}{\partial \rho} = \frac{\partial I(\rho)}{\partial \rho}= \int_{-\infty}^{\infty} \int_{-\infty}^{\infty} 4 \delta(y_1) \delta(y_2) p(y_1, y_2) \, dy_1 dy_2=\frac{2}{\pi \sqrt{1-\rho^2}} }[/math].

When [math]\displaystyle{ \rho=0 }[/math], [math]\displaystyle{ I(\rho)=0 }[/math]. Thus

[math]\displaystyle{ E \left(\text{sign}(y_1) \text{sign}(y_2) \right) = I(\rho)=\frac{2}{\pi} \int_{0}^{\rho} \frac{1}{\sqrt{1-\rho^2}} \, d\rho=\frac{2}{\pi} \arcsin(\rho) }[/math],

which is Van Vleck's well-known result of "Arcsine law".

[2][3]

Application

This theorem implies that a simplified correlator can be designed.[clarification needed] Instead of having to multiply two signals, the cross-correlation problem reduces to the gating[clarification needed] of one signal with another.[citation needed]

References

  1. 1.0 1.1 J.J. Bussgang,"Cross-correlation function of amplitude-distorted Gaussian signals", Res. Lab. Elec., Mas. Inst. Technol., Cambridge MA, Tech. Rep. 216, March 1952.
  2. 2.0 2.1 2.2 Vleck, J. H. Van. "The Spectrum of Clipped Noise". Radio Research Laboratory Report of Harvard University No. 51. 
  3. 3.0 3.1 3.2 Vleck, J. H. Van; Middleton, D. (January 1966). "The spectrum of clipped noise". Proceedings of the IEEE 54 (1): 2–19. doi:10.1109/PROC.1966.4567. ISSN 1558-2256. https://ieeexplore.ieee.org/document/1446497. 
  4. 4.0 4.1 4.2 Price, R. (June 1958). "A useful theorem for nonlinear devices having Gaussian inputs". IRE Transactions on Information Theory 4 (2): 69–72. doi:10.1109/TIT.1958.1057444. ISSN 2168-2712. https://ieeexplore.ieee.org/document/1057444/;jsessionid=p7xQfWaG1zLvg43lhpnzWz6pUrVRPwQvTk_5Z-KclUPBlln2I6MR!144025597. 
  5. 5.0 5.1 Papoulis, Athanasios (2002). Probability, Random Variables, and Stochastic Processes. McGraw-Hill. pp. 396. ISBN 0-07-366011-6. 

Further reading