Intro to Stochastic Systems (Spring 16) Lecture 6

Similar documents
pickup from external sources unwanted feedback RF interference from system or elsewhere, power supply fluctuations ground currents

Fundamentals of Noise

ES 272 Assignment #2. in,3

Conventional Paper I-2010

7.7 The Schottky Formula for Shot Noise

2A1H Time-Frequency Analysis II

Analysis and Design of Analog Integrated Circuits Lecture 14. Noise Spectral Analysis for Circuit Elements

ECE-340, Spring 2015 Review Questions

2A1H Time-Frequency Analysis II Bugs/queries to HT 2011 For hints and answers visit dwm/courses/2tf

Scanned by CamScanner

Laboratory I: Impedance

Consider a parallel LCR circuit maintained at a temperature T >> ( k b

Physics 116A Notes Fall 2004

Lecture 14: Electrical Noise

Electronics and Communication Exercise 1

Electrical Noise under the Fluctuation-Dissipation framework

Transistor Noise Lecture 10 High Speed Devices

Sinusoidal Steady-State Analysis

Prof. Anyes Taffard. Physics 120/220. Voltage Divider Capacitor RC circuits

Handout 11: AC circuit. AC generator

Lecture 23: NorCal 40A Power Amplifier. Thermal Modeling.

Source-Free RC Circuit

Q. 1 Q. 25 carry one mark each.

Lecture 9. PMTs and Laser Noise. Lecture 9. Photon Counting. Photomultiplier Tubes (PMTs) Laser Phase Noise. Relative Intensity

MASSACHUSETTS INSTITUTE OF TECHNOLOGY Physics Department Statistical Physics I Spring Term 2013

Physics GRE: Electromagnetism. G. J. Loges 1. University of Rochester Dept. of Physics & Astronomy. xkcd.com/567/

Physical Noise Sources

Some of the different forms of a signal, obtained by transformations, are shown in the figure. jwt e z. jwt z e

Electronics. Basics & Applications. group talk Daniel Biesinger

QUESTION BANK SUBJECT: NETWORK ANALYSIS (10ES34)

Figure Circuit for Question 1. Figure Circuit for Question 2

EE 40: Introduction to Microelectronic Circuits Spring 2008: Midterm 2

Circuit Analysis-III. Circuit Analysis-II Lecture # 3 Friday 06 th April, 18

Lecture 7: Laplace Transform and Its Applications Dr.-Ing. Sudchai Boonto

Paper Review. Special Topics in Optical Engineering II (15/1) Minkyu Kim. IEEE Journal of Quantum Electronics, Feb 1985

Electric Current. Note: Current has polarity. EECS 42, Spring 2005 Week 2a 1

Mixed Signal IC Design Notes set 6: Mathematics of Electrical Noise

Lecture 24. Impedance of AC Circuits.

Linear Filters. L[e iωt ] = 2π ĥ(ω)eiωt. Proof: Let L[e iωt ] = ẽ ω (t). Because L is time-invariant, we have that. L[e iω(t a) ] = ẽ ω (t a).

Math Fall Linear Filters

0 t < 0 1 t 1. u(t) =

ECE Branch GATE Paper The order of the differential equation + + = is (A) 1 (B) 2

Most matter is electrically neutral; its atoms and molecules have the same number of electrons as protons.

Modeling of noise by Legendre polynomial expansion of the Boltzmann equation

LAB MANUAL EXPERIMENT NO. 7

EE 330 Lecture 22. Small Signal Modelling Operating Points for Amplifier Applications Amplification with Transistor Circuits

Laboratory III: Operational Amplifiers

Exercise: 4. 1 converges. 1. The series. (D) e. (A) 2 ln 2 (C) 2 (B) 2. equal to. 4. If. a and b equation. d x 2. (D) ae 2t. (A) ae t.

P441 Analytical Mechanics - I. RLC Circuits. c Alex R. Dzierba. In this note we discuss electrical oscillating circuits: undamped, damped and driven.

13.42 READING 6: SPECTRUM OF A RANDOM PROCESS 1. STATIONARY AND ERGODIC RANDOM PROCESSES

Introduction to AC Circuits (Capacitors and Inductors)

Name of the Student: Problems on Discrete & Continuous R.Vs

ECE2262 Electric Circuit

An example of answers for the final report of Electronics"

Electric Circuit Theory

General Information. Vishay Semiconductors. Explanation of Technical Data. Type Designation Code for LEDs

6.776 High Speed Communication Circuits Lecture 10 Noise Modeling in Amplifiers

Shot Noise in pn Junction Diodes and Transistors

Whereas the diode was a 1-junction device, the transistor contains two junctions. This leads to two possibilities:

Lecture 3: Signal and Noise

Notes for course EE1.1 Circuit Analysis TOPIC 10 2-PORT CIRCUITS

Sinusoidal Steady State Analysis (AC Analysis) Part II

Energy Storage Elements: Capacitors and Inductors

Chapter 33. Alternating Current Circuits

Revision of Lecture 4

Sinusoids and Phasors

Chapter 5. BJT AC Analysis

Phys 4061 Lecture Thirteen Photodetectors

Lecture 05 Power in AC circuit

= 4. e t/a dt (2) = 4ae t/a. = 4a a = 1 4. (4) + a 2 e +j2πft 2

7 The Waveform Channel

Mod. Sim. Dyn. Sys. Amplifiers page 1

Problems on Discrete & Continuous R.Vs

CIRCUIT ELEMENT: CAPACITOR

GENERALIZATION OF CAMPBELL S THEOREM TO NONSTATIONARY NOISE. Leon Cohen

GATE 2009 Electronics and Communication Engineering

Lecture #3. Review: Power

Resonant Matching Networks

EE292: Fundamentals of ECE

Lecture 15. Theory of random processes Part III: Poisson random processes. Harrison H. Barrett University of Arizona

Name of the Student: Problems on Discrete & Continuous R.Vs

Continuous-Time Frequency Response (II) Lecture 28: EECS 20 N April 2, Laurent El Ghaoui

Chapter 10: Sinusoids and Phasors

Linear Electrical and Acoustic Systems Some Basic Concepts

EE-202 Exam III April 6, 2017

Notes on Electric Circuits (Dr. Ramakant Srivastava)

Mod. Sim. Dyn. Sys. Amplifiers page 1

Chapter 11 AC and DC Equivalent Circuit Modeling of the Discontinuous Conduction Mode

Quantum Light-Matter Interactions

Transistor Noise Lecture 14, High Speed Devices

E40M. RC Circuits and Impedance. M. Horowitz, J. Plummer, R. Howe

E40M Review - Part 1

First and Second Order Circuits. Claudio Talarico, Gonzaga University Spring 2015

ECE2262 Electric Circuits. Chapter 6: Capacitance and Inductance

sections June 11, 2009

A two-port network is an electrical network with two separate ports

OPTI510R: Photonics. Khanh Kieu College of Optical Sciences, University of Arizona Meinel building R.626

Driven RLC Circuits Challenge Problem Solutions

Electric Circuits II Sinusoidal Steady State Analysis. Dr. Firas Obeidat

Electric Circuit Theory

Transcription:

6 Noise 6.1 Shot noise Without getting too much into the underlying device physics, shot noise refers to random current fluctuations in electronic devices due to discreteness of charge carriers. The first analysis of shot noise was published by Walter Schottky in 1918 in the context of vacuum tubes, although shot noise also manifests itself in semiconductor devices, such as diodes and transistors, and in optoelectronic devices as a consequence of the fact that electromagnetic radiation is carried by photons. A vacuum tube is a device that has a pair of terminals or electrodes enclosed in an evacuated glass container. It is used to control the flow of current in a circuit. For example, the simplest type of a vacuum tube, the diode, allows the current to flow in one direction only. These days, there are other types of diodes based on semiconductors, but historically vacuum tubes have been the first active electronic devices. In a diode, electrons are emitted at the cathode and collected at the anode. The duration of each emission event is assumed to be so short that it can be approximated by a unit impulse, and emission events in nonoverlapping time intervals are assumed to be statistically independent. Thus the emission times can be thought of as arrival times of a Poisson process with rate λ, where λ is the average number of electron emissions per unit time. Thus, the current through the diode has a dc component equal to qλ, where q 1.60 10 19 Coulombs is the charge of the electron. Shot noise describes the fluctuations around this dc component. The transit of each electron from the cathode to the anode induces a time-varying current, whose shape depends on the device characteristics. Thus, the current through the diode at time t 0 is described by the stochastic signal I (I t ) t 0 with I t q h(t T k ), (6.1) k1 where T (T k ) k N are the arrival times of a Poisson process with rate λ, and qh(t) is the current induced in the diode by an electron emitted at time t 0. We assume that h(t) 0 for t < 0 (no current before an emission) and that h(t)dt h(t)dt 1 (conservation of charge). 0 The mean and the variance of I t are given by Campbell s theorem: m I (t) E[I t ] qλ (which is the dc component) and σ I (t) Var[I t] q λ h(τ)dτ qλ 0 h (τ)dτ q λ h(τ)dτ qλ 0 h (τ)dτ (which measures the fluctuations due to randomly timed emissions). Denoting by H the Fourier transform of h, we can also write m I (t) qλh(0) and σ I (t) q λ H(0). 1

However, in order to get the full picture of shot-noise current, we will need to compute the autocorrelation of I and its power spectral density. In order to compute R I (t,t τ), we will use the same approach that we relied on to prove Campbell s formula for the variance of I t. Namely, we first write I as the output of an LTI system with impulse response h and with the input Z (Z t ) t 0 given by Z t q δ(t T k ). k1 Here, Z is a train of unit impulses originating at the arrival times T 1,T,... Each impulse corresponds to the emission of a single electron at the cathode. We have already computed the autocorrelation of Z when proving Campbell s theorem: it is given by Therefore, R I (τ) h h R Z (τ) q λ ( q λ R Z (τ) q ( λ λδ(τ) ). (6.) h(τ s)h(s t)r Z (t)ds dt h(τ s)h(s t)ds dt q λ h(τ s)h(s t)δ(t)ds dt h(t)dt) q λ h h(τ) q λ q λ h h(τ), where we have used the definition of the time reversal h(t) h( t), as well as the fact that h(t)dt 1. The power spectral density of I is the Fourier transform of R I: S I (ω) q λδ(ω) q λ H(ω). (6.3) Note the ttwo terms in (6.3): the first term is the dc term, while the second term is due to the filtering effect of h on Z. It is useful to consider an example: Assuming that each electron moves from the cathode to the anode with constant velocity and that a constant time t 0 elapses between the emission and the collection of each electron, it can be shown 1 that the impulse response h has the form 1 t, 0 t t h(t) 0 0 0, otherwise. 000. 1 See, e.g., Section 4.1 in Hermann Haus, Electromagnetic Noise and Quantum Optical Measurements, Springer, Berlin,

Taking the Fourier transform, we get Substituting this into (6.3), we get 6. Johnson Nyquist noise H(ω) h(t)e iωt dt 1 t0 e iωt dt t 0 0 1 e iωt t 0 iωt 0 0 1 e iωt 0 iωt 0 e iωt 0/ e iωt0/ e iω0t/ ωt 0 i e sin( ) ωt 0 iωt 0/ e iωt 0/ sinc ωt 0 ( ωt0 π ). S I (ω) q λδ(ω) q λ sinc ( ωt0 π Shot noise depends only on the current, and is due entirely to the discreteness of charge carriers. Another source of noise in devices is due to thermal agitation of charge carriers, and its effect increases with resistance and with temperature. Let us first consider the case of a noisy resistor with resistance R at absolute temperature T. Using an argument based on the so-called equipartition theorem from statistical physics, Nyquist showed that such a noisy resistor can be described by a Thévenin equivalent model consisting of a noiseless resistor with resistance R in series with a noisy voltage source whose voltage is given by a Gaussian stochastic signal E (E t ) t R with zero mean and the white-noise autocorrelation function R E (τ) kt Rδ(τ). Here, T is the temperature in Kelvin (e.g., 90 K corresponds to room temperature), and k 1.3806 10 3 J/K is the Boltzmann constant. Alternatively, we can use the Norton equivalent, where the noiseless resistor R is connected in parallel with a noisy current source J (J t ) t R given by J t E t R, so that R J(τ) kt Gδ(τ), where G 1/R is the conductance. Consequently, the Nyquist voltage source has the power spectral density S E (ω) kt R, while the Nyquist current source has S J (ω) kt G. Of course, ideal white noise is unphysical (recall that it has infnite average power), and it is often sufficient to assume that the noisy voltage (or current) power spectral density is constant and nonzero only over some finite bandwidth. In that case, we will have kt R, ω πb S E (ω) (6.4) 0, otherwise ). 3

E t I t a R C V t b Figure 1: An RC circuit with a noisy resistor. where B is the bandwidth in Hz, and so E[E t ] 1 π 1 π S E (ω)dω πb 4kT RB. kt Rdω πb Let us consider an example of a circuit with such a noisy resistor. Consider the RC circuit shown in Figure 1. Since E t is a noisy voltage source, all currents and voltages in this circuit are stochastic signals. Let us compute the power spectral density of the voltage across the terminals a and b, i.e., the voltage across the capacitor. Kicrhhoff s voltage law gives I t R E t V t 0, (6.5) where the current through the capacitor is given by I t C dv t dt. Consequently, RC dv t dt V t E t, which means that the transfer function from E to V is given by H(ω) Thus, the power spectral density of V is given by 1 1 iωrc. S V (ω) H(ω) S E (ω) S E (ω) 1 ω (RC). (6.6) In particular, if E is the ideal Johnson Nyquist noisy voltage source, then (6.6) gives S V (ω) kt R 1 ω (RC) ; 4

if the white-noise model is assumed only over a finite bandwidth of B Hz, then kt R S V (ω) 1 ω, if ω πb (RC) 0, otherwise. Another, equivalent, way to express (6.6) is as follows: Let Z(ω) denote the impedance of the circuit between the terminals a and b. Viewed from these terminals, the resistor R and the capacitor C are connected in parallel, and therefore or, equivalently, 1 Z(ω) 1 Z R (ω) 1 Z C (ω) 1 R iωc, Z(ω) R 1 iωrc. Then, for the ideal Johnson Nyquist noise source, we have whereas for the bandlimited one we have S V (ω) kt ReZ(ω), (6.7) S V (ω) 4kT BReZ(ω). (6.8) We will now show that Eqs. (6.7) and (6.8) are special cases of a more general theorem of Nyquist: The voltage across any pair of terminals a, b in a network consisting of noisy resistors at temperature T, inductors, and capacitors has the power spectral density for ideal Johnson Nyquist noise and S V (ω) kt ReZ(ω) (6.9) S V (ω)4kt BReZ(ω) (6.10) for bandlimited Johnson Nyquist noise with bandwidth B, where Z(ω) is the impedance of the network across the termimal pair a,b. In order to prove Nyquist s theorem, we first need to introduce some background concepts. A two-port network, shown in Figure, is a circuit with two ports, i.e., two pairs of terminals that can be connected to external circuits. Each port has a voltage and a current, and the relationship between the voltages and the currents is specified by the so-called transfer matrix ( ) ( )( ) V1 A B V, C D I 1 where V 1 is the voltage across port 1, I 1 is the current into port 1, V is the voltage across port, and I is the current out of port. It is important to keep in mind that this is a frequency-domain description. We say that the two-port circuit is: 5 I

I 1 I A B V 1 T V C D Figure : A two-port circuit. reciprocal if the transfer function from I 1 to V is equal to the transfer function from I to V 1 : V (ω) I 1 (ω) V 1(ω) I (ω), ω (the minus sign is there because I is the current flowing out of port ); lossless if the power supplied at port 1 is equal to the power that exits at port, and vice versa. A network that consists only of resistors, inductors, and capacitors is reciprocal; a network that consists only of inductors and capacitors is lossless. We will assume, for simplicity, that our network has only one resistor, and all other elements are lossless. If we represent the noisy resistor by its Norton equivalent consisting of a noiseless resistor in parallel with a noisy current source J (J t ) t R, then we can view J as a (noisy) current I 1 (I 1,t ) into port 1 of a two-port network and the voltage across the resistor as the noisy voltage V 1 (V 1,t ) across port 1, and where we associate port with the terminals a and b see Figure 3. Thus, the voltage signal of interest is the voltage V (V,t ) across port. The power flowing into port is equal to I (ω) ReZ(ω), and, since the I 1 I V A B 1 R T V C D Figure 3: A general passive network with a noisy resistor; the T -matrix represents the rest of the network, which consists only of inductors and capacitors. network connected to the resistor is lossless by assumption, all of this power is dissipated at the resistor. That is, I (ω) ReZ(ω) V 1(ω). (6.11) R 6

Let H denote the transfer function from I 1 to V. By reciprocity, it is equal to the transfer function from I to V 1. Using this fact together with (6.11), we can write H(ω) V (ω) I 1 (ω) V 1 (ω) I (ω) Since the power spectral density of the voltage V V is given by S V (ω) H(ω) S J (ω), RReZ(ω). (6.1) we can use (6.1) and the fact that S J (ω) S E(ω) R to obtain Eqs. (6.9) and (6.10). 6.3 Amplifier noise A good illustration of the theory of noise sources in circuits is the analysis of amplifier noise. An amplifier is an active device that provides a power gain relative to its input. A noisy amplifier, just like its name suggests, also introduces additional noise at the output. In order to discuss amplifiers, we first must introduce the notions of a noisy source and exchangeable power. A source, shown in Figure 4 (left), is a one-port circuit (i.e., a circuit with a single pair of terminals) that has the voltage-current characteristic V Z S I E, (6.13) where V is the voltage across the source s terminals, Z S is the source impedance, I is the current flowing into the source, and E is the open-circuit voltage of the source. We say that the source is noisy if its open-circuit voltage E is a stochastic signal. For example, if the source is a noisy resistor with resistance R at temperature T, then we can take E to be the bandlimited Johnson Nyquist voltage noise. I E I E V Z S Z L V Z S Figure 4: The equivalent circuit of a source (left) and a source connected to a load (right). Suppose that we connect the source to a load with load impedance Z L, see Figure 4 (right). The average power delivered to the load at time t is given by P L,t E[ I t ]ReZ L, 7

where the Fourier transform of I t can be computed from the Kirchhoff s voltage law: Therefore, I(ω) E(ω) Z L Z S. P L,t E[ E t ]ReZ L Z S Z L. (6.14) If E is weakly stationary, then E[ Et ] does not depend on t, and we can express (6.14) as P L R E(0)ReZ L Z S Z L, (6.15) where R E (τ) denotes the autocorrelation function of E. The quantity in (6.15) is a function of the load impedance Z L. Exchangeable power, denoted by P ex, is the global extremum (maximum or minimum) value of the power fed to the load by the source. Exchangeable power can be positive (the source dissipating power into the load) or negative (the load dissipating power into the source). To compute it, we first separate the load and the source impedances into the resistive and the reactive components: Z L R L ix L and Z S R S ix S. Then and since P L R E (0)R L (R L R S ) (X L X S ), P L R E (0)(X L X S ) X L ((R L R S ) (X L X S ) ), we see that X L X S is a necessary condition for extremum. Therefore, we assume that X L X S, and so now we need to compute the extremum of R E (0)R L (R L R S ) over R L. In this case, it is a simple exercise in calculus to show that, upon setting R L R S, the power delivered to the load achieves its global maximum if R S > 0 and global minimum if R L < 0. In either case, the global extremum value is the exchangeable power, given by P ex R E(0) (Z S ZS (6.16) ), and it is achieved by matching the load impedance to the source impedance, i.e., Z L Z S. Now suppose that we connect the source to one port of a two-port circuit with transfer matrix T, as shown in Figure 5. The exchangeable gain G of the resulting network is defined as the ratio of 8

Z S E I 0 1 I A B V 0 T C D 1 V Figure 5: A noisy source connected to a noiseless two-port circuit. the output exchangeable power P ex, (i.e, at port ) to the input exchangeable power P ex,1 (i.e., at port 1). The latter is given by (6.16): P ex,1 R E(0) (Z S Z S ). To compute P ex,, we need a voltage-current characteristic like (6.13) to relate V, I, and E. To that end, we first write down the relation between the input and output currents and voltages for our two-port: V 1 AV BI, I 1 CV DI. (6.17a) (6.17b) Substituting the expressions (6.17) into Kirchhoff s voltage law V 1 E I 1 Z S, we obtain Rearranging, we get AV BI (CV DI )Z S E. V B DZ S 1 I A CZ E, S A CZ S so, looking from port, we can view the whole circuit as a source with effective impedance Z S BDZ S E ACZ and open-circuit voltage Ẽ S ACZ. Therefore, the output exchangeable power is given S by Now, since P ex, RẼ(0) ( Z S Z S (6.18) ). R Ẽ (0) R E (0) A CZ S 9

and we get from (6.18) Z S Z S B DZ S A CZ S B D Z S A C Z S P ex, and can now write down the exchangeable gain: (B DZ S)(A CZ S ) (B DZ S ) (A CZ S ) A CZ S, R E (0) [(B DZ S )(A CZ S ) (B DZ S ) (A CZ S )], G P ex, Z S ZS P ex,1 (B DZ S )(A CZ S ) (B DZ S ) (A CZ S ) (6.19) This expression looks rather formidable, but it can be written down more succinctly using matrix notation. Define the following complex vector v and matrix S: ( ) ( ) v 1, S 0 1, 1 0 Z S and recall that the Hermitian conjugate A of a matrix A is obtained by taking the complex conjugate of each entry, followed by taking the transpose. Then we can write (prove this!). An amplifier has G 1. G v Sv v T ST v (6.0) Z S I 1 E V 1 E 0 J 0 noisy network I 0 1 I A B V 0 T C D 1 V noisy amplifier noise-free network Figure 6: An equivalent circuit model of a noisy amplifier. Now we are going to consider the effect of internal noise in the amplifier. Amplifier noise is typically modeled by including internal noisy voltage and current sources into the amplifier 10

network, as shown in Figure 6. We assume that the noisy voltage E 0 and the noisy current J 0 are independent of the noisy source voltage E. The effect of these internal noise sources manifests itself at the output of the amplifier. To quantify it, one typically uses the so-called noise figure F, which is defined as the ratio of the output signal-to-noise ratio to the input signal-to-noise ratio. In our case, the noise figure is given by the ratio of the variance of the voltage V 1 at the input to the noiseless part of the amplifier and the variance of the source voltage E. To compute the variance of V 1,t, we first replace the combination of the source and the noisy amplifier network by its Thévenin equivalent circuit with source impedance Z S in series with the Thévenin voltage source E Th E E 0 JZ S. Then V 1 ETh, and, since E is independent of E 0 and J 0, E[ V 1,t ] E[ E t ] E[ E 0,t J 0,t Z S ]. Assuming that all noise sources are jointly weakly stationary, we have E[ E t ] R E (0) and E[ E 0,t J 0,t Z S ] E[ E 0,t ] ReE[E 0,t J 0,t] E[ J 0,t ] Therefore, the amplifier noise figure is given by F E[ V 1,t ] E[ E t ] R E0 (0) Re,R E0 J 0 (0) R J0 (0). R E(0) R E0 (0) Re,R E0 J 0 (0) R J0 (0) R E (0) 1 R E 0 (0) Re,R E0 J 0 (0) R J0 (0). R E (0) Note that F 1. If the input source consists of an impedance Z S in series with a bandlimited Johnson Nyquist noise source with bandwidth B at temperature T, then R E (0) 4kT BReZ S, and therefore we can write down the so-called excess noise figure F 1 R E 0 (0) Re,R E0 J 0 (0) R J0 (0) 4kT BReZ S R Th (0) kt B(Z S ZS ), where R Th is the autocorrelation function of E Th. Thus, F 1 is equal to the exchangeable power P Th ex at the source with impedance Z S and noisy open-circuit voltage E Th, divided by kt B: F 1 P ex Th kt B. Now that we have obtained the expressions for the exchangeable gain G and the noise figure F, we 11

can compute the so-called amplifier noise measure: For high-gain amplifiers (G 1), M F 1. 6.4 1/f noise M F 1 1 1/G Pex Th kt B. v Sv v T ST v Both shot noise and the Johnson Nyquist noise are due to fundamental physical phenomena, namely discreteness of charge carriers and their thermal agitation. These two effects would still be present in ideal devices. However, measurements reveal another noise phenomenon due to the presence of impurities. This phenomenon, referred to as 1/f noise for reasons that will become clear shortly, manifests itself as random fluctuations of the conductance of the device due to spontaneous generation and recombination of electrons and holes. These fluctuations are especially noticeable at low frequencies, which is why frequency upconversion is often employed in low-noise electronic devices. In fact, experiments have shown that the power spectrum of these conductance fluctuations behaves as 1/f, where f ω/π is the frequency in Hertz. This is the reason for the name 1/f noise. As we will now see, a simple model of 1/f noise can be derived by starting from a toy Markovchain model of generation-recombination noise in a nearly pure semiconductor. Consider a continuous-time stochastic signal X (X t ) t R with binary state space X { 1,1}. The state 1 corresponds to the spontaneous generation of an electron, the state 1 corresponds to the spontaneous generation of a hole, and we will arrange things in such a way that, on average, there are no charge fluctuations, so P[X t 1] P[X t 1] 1. For s t, let M s,t(, ) denote the conditional probability that the X t 1 given X s 1, and define M s,t (,), M s,t (, ), and M s,t (,) analogously. We will first construct a time-discretized approximation, and then look at the continuous-time limit. Thus, pick a sufficiently small value τ > 0 and assume that the state can change at times..., τ, τ,0,τ,τ,..., and that the state transitions at different times are independent. Let p t ( ) P[X t 1]. We will assume that M t,th (,) M t,th (, ) αh for some α > 0, i.e., α is the rate of state transitions per unit time. Then, for any t R and any h < 1/α, we have and an analogous derivation gives p th ( ) p t ( )M t,th (, ) p t ()M t,th (, ) (1 αh)p t ( ) αhp t (), p th () αp t ( ) (1 α)p t (). In matrix form, we obtain ( pth ( ) p th () ) ( p t ( ) p t () )( ) (1 α)h αh, (6.1) αh 1 αh 1

which, upon rearranging, becomes 1 ( pth ( ) p h t ( ) p th () p t () ) ( p t ( ) p t () )( ) α α. α α Taking the continuous-time limit h 0, we end up the first-order matrix ODE d ( pt ( ) p dt t () ) ( p t ( ) p t () )( ) α α. (6.) α α It can be solved by diagonalizing the matrix on the right, and, for any t R and any τ 0, the solution is given by ( ptτ ( ) p tτ () ) ( p t ( ) p t () ) 1 e ατ 1 e ατ 1 e ατ 1 e ατ. (6.3) For s t, define the matrix ( ) 1 e α(t s) Ms,t (, ) M M s,t s,t (,) M s,t (, ) M s,t (,) 1 e α(t s) 1 e α(t s) 1 e α(t s). (6.4) Then M s,s I, p t p s M s,t, and a simple calculation shows that the probability vector π (1/,1/) T is invariant under all M s,t, i.e., if p s ( ) p s () 1/, then p t ( ) p t () 1/ as well. We will assume, therefore, that p t ( ) p t () 1/ for all t, corresponding to no excess electrons or holes on average. Since and P[X t X tτ ] P[X t 1,X tτ 1] P[X t 1,X tτ 1] p t ( )M t,tτ (, ) p t ()M t,tτ (,) 1 e ατ P[X t X tτ ] P[X t 1,X tτ 1] P[X t 1,X tτ 1] p t ( )M t,tτ (,) p t ()M t,tτ (, ) 1 e ατ, the autocorrelation function of X is given by R X (t,t τ) E[X t X tτ ] P[X t X tτ ] P[X t X tτ ] e ατ. 13

A similar calculation shows that R X (t τ,t) e ατ, so we obtain R X (τ) e α τ. (6.5) It is convenient to define the relaxation time t 0 1/α and express (6.5) as R X (τ) e τ /t 0. Now let us compute the power spectral density of X: S X (ω) R X (τ)e iωτ dτ e τ /t 0 e iωτ dτ 0 e (1/t0 iω)τ dτ 1 1/t 0 iω 1 1/t 0 iω /t 0 1/t0 ω t 0 1 (t 0 ω). 0 e (1/t 0iω)τ dτ In this context, it is customary to work with the frequency in Hertz, f ω/π, so we obtain S X (f ) t 0 1 (πt 0 f ), (6.6) which is known as the Lorentzian power spectrum, see Figure 7. The Lorentzian power spectrum describes the conductance fluctuations in a nearly pure semiconductor, where t 0 is the characteristic timescale of electron-hole recombination. However, in a typical semiconductor device with impurities, there are multiple generation-recombination processes going on, each with its own relaxation time. A good physical model, as it turns out, results if we assume that the relaxation time is a nonnegative random variable with a pdf g, such that, for any t t 0 t and any n times t 1 < t <... < t n, P[t T 0 t,x t1 x 1,...,X tn x n ] 1 Under this assumption, and t n 1 P[X t X tτ ] P[X t τ X t ] 1 P[X t X tτ ] P[X t τ X t ] 1 14 t 1 ( 1) x kx k1 e (t k1 t k )/t g(t)dt. k1 g(t)(1 e τ/t )dt g(t)(1 e τ/t )dt,

S X (f).0 1.5 1.0 0.5 - -1 1 f Figure 7: The Lorentzian power spectrum, shown here for t 0 1. which gives the autocorrelation function and the power spectrum R X (τ) g(t)e τ /t dt tg(t) S X (f ) dt. (6.7) 1 (πtf ) We will now show that the 1/f power spectrum results if we choose the pdf g appropriately. From semiconductor physics, we know that, in thermal equilibrium at absoulte temperature T, the relaxation time in a material with the energy difference E between the electron and the hole energy levels is given by ce E/kT for some c > 0, where k is the Boltzmann constant. If we assume that the energy gap E is a random variable with a uniform distribution over some range [ 0, 1 ], then the pdf g of the relaxation time will be given by 1 g(t) kt ( 1 0 )t, for ce 0/kT t ce 1/kT. (6.8) 0, otherwise Substituting (6.8) into (6.7), we get ce 1 /kt 1 S X (f ) kt ( 1 0 ) 1 (πtf ) dt 1 πf kt ( 1 0 ) tan 1 (πf t) ce 1 /kt which behaves like 1/f for large values of 1 0. ce 0 /kt Last version: April 5, 016 ce 0 /kt, 15