yamamoto noise

Upload: boris-berkopec

Post on 06-Apr-2018

236 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/3/2019 Yamamoto Noise

    1/195

    1. Introduction

    When one tries to measure a weak signal, a lower limit set by the spontaneous fluctuation in

    the current, voltage, and other physical variables of the system under test is always encountered.

    These spontaneous fluctuations are referred to as noise. Noise is a significant problem in science

    and engineering because it places an ultimate sensitivity on all measurements. In this class, the

    mathematical methods, physical origins and characteristics of noise in various devices and systems

    will be studied.

    Chapter 2 provides a mathematical method for treating noise. Two complimentary averaging

    methods, time averaging and ensemble averaging, and the experimental realization of the two

    averaging processes are described. The Fourier analysis, specifically the Wiener-Khintchin theorem,

    is introduced to calculate the power spectrum and autocorrelation function of a noisy waveform.

    A noisy waveform often consists of discrete random pulse trains. Various statistical distribution

    functions of random pulse trains (i.e., binomial, Poisson, sub-Poisson, super-Poisson, and Gaussian)

    are derived and the effect of random deletion of pulses (partition noise) is discussed using the

    Burgess variance theorem. A method for calculating the power spectrum and autocorrelation

    function of random pulse trains using the Carson theorem is presented.

    Chapter 3 describes how to calculate the noise figure of linear networks. A noisy electrical

    network can be represented by a noise-free network with external noise generators. If such a network

    has two ports (input and output), the noise figure is often used as a figure of merit expressing the

    inherent noisiness of the circuit. The technique is also useful to express the noise properties of

    optical systems such as laser and parametric amplifier systems.

    Chapter 4 discusses the physical origins of two types of intrinsic noise: thermal noise and

    1

  • 8/3/2019 Yamamoto Noise

    2/195

    quantum noise. Thermal noise of a resistor is calculated by microscopic theory using a Brownian

    particle model or by macroscopic theory using the equipartition theorem. The latter approach,

    developed by Nyquist, is very general and easily extended to include quantum noise (i.e., the

    effect of zero-point fluctuation). The expression for the power spectrum of an open circuit voltage

    fluctuation, including both thermal and quantum noise, is called a generalized Nyquist formula.

    The noise characteristics of a mesoscopic conductor on an extremely small scale is enlightening

    with regard to the physical origins of thermal and quantum noise. The transition from mesoscopic

    conductor to macroscopic conductor will be discussed in some detail.

    Chapter 5 treats the noise of semiconductor pn junctions diodes. Random diffusive motion

    and generation-recombination event of carriers contribute to the noise of a pn junction. A pn-

    junction diode driven by a low-impedance, constant-voltage source has a shot-noise-limited current

    fluctuation, but a junction voltage fluctuation is suppressed. A pn-junction diode driven by a high-

    impedance, constant-current source has a shot-noise-limited voltage fluctuation, but a junction

    current fluctuation is suppressed. A collective and single charge Coulomb blockade effect in a pn

    junction is discussed as the principle of sub-Poisson light generation.

    Chapter 6 discusses the noise of a bipolar transistor using the method of a noisy linear network.

    Chapter 7 describes the mathematical models and physical mechanisms of 1/f noise. In prac-

    tically all electronic and optical devices, the excess noise obeying the inverse frequency power law

    exists in addition to intrinsic thermal and quantum noise. This 1/f noise places a serious limit on

    the sensitivity of precision measurements at a very low frequency.

    Chapter 8 treats the noise of a tunnel diode. While a macroscopic tunnel diode driven by

    a constant voltage source has a shot-noise-limited current fluctuation, a mesoscopic tunnel diode

    2

  • 8/3/2019 Yamamoto Noise

    3/195

    features various sub-shot-noise behaviors due to single charge Coulomb blockade effect.

    Chapter 9 describes the noise of negative conductance oscillators which generate a coherent

    electromagnetic field at various frequencies. A laser is also described as a negative conductance

    oscillator. A negative conductance oscillator consists of a frequency selective element which provides

    positive feedback and a gain element which provides signal amplification and nonlinearity. A general

    electrical circuit theory (van der Pol oscillator model) is presented to calculate the amplitude and

    phase noise of the coherent radiation. The difference between the cavity internal field noise and

    output field noise, and the difference between a free-running oscillator and an injection-locked

    oscillator, are pointed out.

    Chapter 10 provides a brief overview of the noise of parametric amplifiers. The fundamental

    limit of phase-insensitive and phase-sensitive linear amplifiers are discussed. A degenerate para-

    metric amplifier is introduced as a noise-free, phase-sensitive amplifier.

    Chapter 11 provides a brief overview of the noise in optical communication systems. The

    signal-to-noise ratio and bit error rate of optical-preamplifier and repeater-amplifier systems are

    calculated and compared with a direct detection scheme. The concept of signal regeneration for

    a digital system and the advantage of using optical amplifiers are discussed. The thermal and

    quantum limits of communication systems are elucidated using the channel capacity.

    3

  • 8/3/2019 Yamamoto Noise

    4/195

    2. Mathematical Methods

    2.1 Time Averaging vs. Ensemble Averaging

    Noise is a stochastic process of a randomly varying function of time and thus is only statis-

    tically characterized. One cannot argue a single event at a certain time; one can only discuss the

    averaged quantity over a certain time interval (time average) or many identical systems (ensemble

    average). Let us consider N identical systems which produce noisy waveforms x(i)(t), as shown in

    Fig. 2-1.

    x(1)(t)

    x(2)(t)

    x(N)

    (t)

    time average

    ensemble average

    t

    t

    t

    Figure 2-1

    One can define the following time-averaged quantities for the i-th member of the ensemble:

    x(i)(t) = limT

    1

    T T

    2

    T2

    x(i)(t)dt ,

    (mean = first-order time average) (2.1)

    1

  • 8/3/2019 Yamamoto Noise

    5/195

    x(i)(t)2 = lim

    T

    1

    T

    T2

    T2

    x(i)(t)

    2dt ,

    (mean square = second-order time average) (2.2)

    (i)x () x(i)(t)x(i)(t + ) = limT

    1

    T

    T2

    T2

    x(i)(t)x(i)(t + )dt .

    (autocorrelation function) (2.3)

    These time-averaged quantities directly correspond to analog measurements using a simple inte-

    grator (low-pass filter) for the measurement of the mean (2.1), square-law detector followed by an

    integrator for the measurement of the mean square (2.2) and delayed correlator followed by an

    integrator for the measurement of the autocorrelation function. Alternatively, one can measure the

    noisy waveform x(i)(t) using an analog-to-digital (A/D) coverter and sampling circuit and then

    calculate these quantities using a computer (digital measurements).

    One can also define the following ensemble-averaged quantities for all members of the ensemble

    at a certain time:

    x(t1) = limN 1N

    Ni=1

    x(i)(t1) =

    x1p1(x1, t1)dx1 ,

    (mean = first-order ensemble average) (2.4)

    x(t1)2 = limN

    1

    N

    Ni=1

    x(i)(t1)

    2=

    x21p1(x1, t1)dx1 ,

    (mean square = second-order ensemble average) (2.5)

    x(t1)x(t2) = limN

    1

    N

    N

    i=1

    x(i)(t1)x(i)(t2) (2.6)

    =

    x1x2p2(x1, t1; x2, t2)dx1dx2 .

    (covariance = second-order joint moment)

    2

  • 8/3/2019 Yamamoto Noise

    6/195

    Here, x1 = x(t1), x2 = x(t2), p1(x1, t1) is the first-order probability density function, and

    p2(x1, t1; x2, t2) is the second-order joint probability density function. p1(x1, t1)dx1 is the prob-

    ability that x is found in the range between x1 and x1 + dx1 at a time t1 and p2(x1, t1; x2, t2)dx1dx2

    is the probability that x is found in the range between x1 and x1 + dx1 at a time t1 and also in the

    range between x2 and x2 + dx2 at a different time t2.

    If x(t1) and x(t1)2 are independent of the time t1 and if x(t1)x(t2) is dependent only on the

    time difference = t2 t1, such a noise process is called a statistically-stationary process. For a

    statistically-nonstationary process, the above is not true. In such a case, the concept of ensemble

    averaging is still valid, but the concept of time averaging fails. An ensemble average is a convenient

    theoretical concept since it is directly related to the probability density functions. On the other

    hand, time averaging is more directly related to real experiments. One cannot prepare an infinite

    number of identical systems in a real situation. Theoretical predictions based on ensemble averaging

    are equivalent to the experimental measurement of time averaging when, and only when, the system

    is a so-called ergodic ensemble. One can loosely say that ensemble averaging and time averaging

    are identical for a statistically-stationary system, but are different for a statistically-nonstationary

    system.

    2.2 Fourier Analysis

    When x(t) is absolutely integrable, i.e.,

    |x(t)|dt < , (2.7)

    3

  • 8/3/2019 Yamamoto Noise

    7/195

    the Fourier transform of x(t) exists and is defined by

    X(j) =

    x(t)ejtdt . (2.8)

    The inverse transform is given by

    x(t) =1

    2

    X(j)ejtd . (2.9)

    This is proven by taking the integral in (2.9) over the limits N (instead of), substituting for

    X(j) from (2.8), and interchanging the order of integration to obtain

    12

    X(j)ejtd = 1

    lim

    N

    x(t) sin{N(t

    t)}(t t) dt

    =

    x(t)(t

    t)dt

    = x(t) . (2.10)

    When x(t) is a real function of time, as it always is the case for an observable waveform, the

    real part of X(j) is an even function of and the imaginary part is an odd function of [i.e.,

    X(j) = X(j)].

    When x(t) is a statistically-stationary process, condition (2.7) is not satisfied and thus the

    Fourier transform does not exist. The total energy of the noisy waveform x(t) is infinite, but

    the noise power (energy flow per second) can be finite. In any practical noise measurement, a

    measurement time interval T is finite and the energy of such a gated function xT(t), defined by

    xT(t) =

    x(t)|t|

    T

    2

    0 |t| > T2, (2.11)

    is also finite. The Fourier transform of such a gated function does exist.

    4

  • 8/3/2019 Yamamoto Noise

    8/195

    If x1(t) and x2(t) have Fourier transforms X1(j) and X2(j), one obtains

    x1(t)x

    2(t)dt =

    1

    2

    X1(j)X

    2 (j)d .

    (Parseval theorem) (2.12)

    This relation is known as the Parseval theorem and is proven by substituting the inverse transform

    ofx2(t) into the LHS of (2.12), interchanging the order of integration, and identifying the resulting

    integral about t as X1(j). If one uses x1(t) = xT(t + ) and x2(t) = xT(t) in (2.12), one obtains

    xT(t + )xT(t)dt =1

    2

    |XT(j)|2eid . (2.13)

    When = 0, (2.13) is reduced to

    [xT(t)]2dt =

    1

    2

    |XT(j)|2d .

    (Energy theorem) (2.14)

    The physical interpretation ofXT(j) and |XT(j)|2 is now clear from the above relations. XT(j)

    is the (complex) amplitude of the ejt component in a gated function xT(t) and |XT(j)|2 is the

    energy density of xT(t) with units of energy per Hz. Equation (2.14) is the total energy of xT(t)

    and increases linearly with T for a statistically-stationary process.

    The average power of xT(t), defined by

    limT

    1

    T

    [xT(t)]2dt = lim

    T

    1

    2

    0

    2|XT(j)|2T

    d , (2.15)

    is a constant and universal quantity. If ensemble averaging is taken for a gated function xT(t) in

    (2.15), the order of limT

    and

    0d can be interchanged. The power spectral density is defined as

    Sx() = limT

    2|XT(j)|2T

    .

    5

  • 8/3/2019 Yamamoto Noise

    9/195

    (unilateral power spectral density) (2.16)

    When a statistically-stationary noisy waveform x(t) is input into an RF spectrum analyzer, the

    above (unilateral) power spectral density is displayed on the screen, usually with units of power

    dBm = 10log10 P(mW), as shown in Fig. 2-2. The noisy waveform x(t) beats against a strong

    local oscillator wave at a frequency in a mixer and only the low-frequency part of the beat

    signal is detected by using a low-pass filter with a bandwidth f, where f is termed a resolution

    bandwidth. The detected noise power is thus due to the frequency component at of the noisy

    waveform x(t). An RF spectrum analyzer measures the noise powers of a fixed frequency for N

    wavepackets with a duration of T = 1/f and displays the averaged noise power. This is essentialy

    the ensemble average procedure. The averaging time T N determines a video bandwidth f /N. If

    one divides this averaged noise power by a resolution bandwidth f, one can obtain the (ensemble

    averaged) power spectral density Sx().

    Local Oscillator wave A sin t

    Noisy waveform x(t)

    f = : resolution bandwidth1

    T

    = : video bandwidth1

    TN

    f =1

    T

    1

    Np(i)i

    TT TTTT

    N

    t

    mixer

    ~

    low-pass filter signal averagert

    noisepower

    frequency

    envelope detector

    Figure 2-2

    6

  • 8/3/2019 Yamamoto Noise

    10/195

    When = 0 in (2.13), one can also divide both sides of (2.13) by T, take an ensemble average,

    and take a limit as T to obtain

    limT

    1T

    xT(t + )xT(t)dt = lim

    T

    12

    0

    2|X(j)|2T

    cos d . (2.17)

    The LHS of this expression is the (ensemble averaged) autocorrelation function x(). Using (2.16)

    in the RHS of this expression, one obtains

    x() =1

    2

    0

    Sx()cos d . (2.18)

    The inverse relation of this expression is

    Sx() = 4

    0

    x()cos d .

    (Wiener-Khintchine theorem) (2.19)

    Equation (2.19) is proven by using (2.18) for x() in the RHS of (2.19), interchanging the order

    of the integral, and identifying the resulting integral about as a Dirac delta function ( ).

    Equations (2.18) and (2.19) constitute the Wiener-Khintchine theorem and indicate that 2 x()

    and Sx() are the Fourier transform pairs.

    If a noisy waveform x(t) is a statistically-stationary process, as shown in Fig. 2-3, with an

    exponentially decaying autocorrelation function

    x() = x(0) exp

    ||

    1

    , (2.20)

    where x(0) = x2 and 1 is a relaxation time constant (systems memory time). Substituting (2.20)

    into (2.19), one obtains the unilateral power spectral density

    Sx() =4x(0)11 + 221

    . (2.21)

    7

  • 8/3/2019 Yamamoto Noise

    11/195

    statistically-stationarynoisy waveform

    statistically-nonstationarynoisy waveform

    x(t)

    t = 0

    y(t) = x(t')dt'0

    t

    t = 0

    y(t)2

    = Dyt

    t

    t

    Figure 2-3

    The spectrum is Lorentzian with a cut-off frequency of 1/1 and the low-frequency spectral density

    is Sx( = 0) = 4x(0)1. The autocorrelation function and the unilateral power spectrum are

    shown in Fig. 2-4.

    x( )S

    x( )4

    01

    x() x( )0

    1

    0.5

    0 1-1 /1 10-2 10-1 101

    10-2

    10-1

    1

    1

    Autocorrelation Function Unilateral Power Spectrum

    Figure 2-4

    A time-integrated function y(t) =t0 x(t

    ) dt

    of a statistically-stationary process x(t

    ) goes

    through a random walk diffusion, as shown in Fig. 2-3. This is called a Wiener-Levy process and

    8

  • 8/3/2019 Yamamoto Noise

    12/195

    is an example of a statistically-nonstationary process. A gated function is defined by

    y(t) =

    t0 x(t

    ) dt (0 t T)

    0 (otherwise)

    . (2.22)

    The corresponding autocorrelation function becomes a function of and T:

    y(, T) =1

    T

    T||0

    y(t + )y(t) dt

    =T

    2

    1 ||

    T

    2Dy . (2.23)

    In order to derive the second line of (2.23), y(t) was used as a cumulative process and thus

    y(t + )y(t) = [y(t) + y(t + )]y(t) = y(t)2 = Dyt, where Dy is a diffusion constant. The unilat-

    eral power spectral density is given by

    Sy(, T) = 4

    T0

    y(, T) cos() d

    =4Dy2

    1 sin(T)

    T

    . (2.24)

    The correlation time is now proportional to the measurement time interval T and a finite mea-

    surement time T prevents the divergence of the power spectral density at = 0, as shown in

    Fig. 2-5.

    2.3 Random Pulse TrainsBinomial, Poisson, and Gaussian Distributions

    A noisy waveform x(t) often consists of a very large number of random, discrete events and

    are represented by

    x(t) =Kk=1

    akf(t tk) . (2.25)

    9

  • 8/3/2019 Yamamoto Noise

    13/195

    0-1 1/

    y(, T) /1

    2DyT

    1

    0.5

    Sx( , T) / 24DyT2

    10-2

    10-1

    1

    10-2 10-1 1 10

    Autocorrelation Function Unilateral Power Spectrum

    Figure 2-5

    One assumes the pulse amplitude ak and the time of event tk are random variables, but that the

    pulse-shape function f(t) is a fixed function, as shown in Fig. 2-6. In a real physical situation, f(t)

    is determined, for example, by the relaxation time of a system or the transit time of a carrier, and

    thus it is a reasonable assumption that f(t) is a fixed function.

    a1

    a2

    aka4

    a3

    t1 t2 tkt4t3t

    measurement time interval T = N

    Figure 2-6

    Suppose the average rate of such random pulse events per second is and the measurement

    time interval T is divided into N = T time slots with a duration . The probability p of pulse

    emission in each time slot is , which, by making N large enough, can be much smaller than one.

    In such a case, the probability of more than two pulses being emitted in the same slot is negligible.

    If a time slot duration is much longer than a systems correlation time (memory time), each pulse

    10

  • 8/3/2019 Yamamoto Noise

    14/195

    emission occurs independently (that is, the pulse emission at a certain time is not influenced

    by the previous events of the system). In such a case, the probability of finding K events in the

    measurement time interval T (out of N trials) is given by the binomial distribution:

    WN(K) =N!

    K!(N K)!pKqNK , (2.26)

    where q = 1 p is the probability of zero pulse emission. Since this is the expansion coefficient of

    (p + q)N, the binomial distribution WN(K) satisfies the normalization condition,

    N

    K=0

    WN(K) = (p + q)N = 1 . (2.27)

    The mean of the number of pulses is calculated by

    KN

    K=0

    KWN(K) =N

    K=0

    N!

    K!(N K)!p

    p(pK)

    qNK

    =

    p

    p

    (p + q)N

    = pN . (2.28)

    The mean-square is

    K2 N

    K=0

    K2WN(K) =N

    K=0

    N!

    K!(N K)!

    p

    p

    2(pK)

    qNK

    =

    p

    p

    2(p + q)N

    = (pN)2 + pqN . (2.29)

    The variance of the number of pulses is then given by

    K2 K2 K2 = N p(1 p) . (2.30)

    11

  • 8/3/2019 Yamamoto Noise

    15/195

    We start with a fixed (constant) number of potential events N = T / and introduce random

    deletion with the probability q = 1 p. The variance of the output event (2.30) is a rather general

    result for such a stochastic process and is referred to as partition noise. In some cases, the

    number of potential events N itself fluctuates. In such a case, the variance of the number of output

    events is given by

    K2 = N2p2 + N p(1 p) ,

    (Burgess variance theorem)

    (2.31)

    where N2 = N2 N2. The above relation is known as the Burgess variance theorem and is

    proven by replacing WN(K) in (2.28) and (2.29) by W(N)WN(K), where W(N) is the probability

    distribution function for potential events N. The first term on the RHS of (2.31) indicates that

    the fluctuation of the initial number of events is suppressed by a loss probability p. The second

    term, on the other hand, indicates the new noise term (partition noise).

    In the limit ofp 1 and N 1, one obtains the following approximate relations:

    (1 p)NK

    ep(NK)

    epN

    , (2.32)

    lnN!

    (N K)!

    d

    dNln N!

    K ln NK . (2.33)

    Using these relations in the binomial distribution (2.26), the approximate expression for WN(K)

    is obtained as

    W(p)N (K) =

    (pN)K

    K!epN =

    KK

    K!eK .

    (Poisson distribution) (2.34)

    This is the Poisson distribution. The variance of the output events K2 can be calculated as

    12

  • 8/3/2019 Yamamoto Noise

    16/195

    follows:

    K2 K

    K=0

    K(K 1)W(p)N (K) = K2

    K=2

    KK2

    (K

    2)!eK = K

    2

    K2 K2 K2 = K = N p . (2.35)

    The result is, of course, obtained directly from (2.30) by letting p 0, but keeping Np finite. The

    same variance is also obtained from (2.31) in the limit p 0. Therefore, the Poisson distribution is

    a quite general distribution for a random-point process. Whenever a random-deletion process with

    a very large attenuation rate is imposed, the final statistics always obey the Poisson distribution,

    irrespective of an initial distribution, as shown in Fig. 2-7.

    However, p cannot be always made much smaller than one for a given average rate of random

    pulse events per second because any physical system has a finite memory time m. If the time

    slot = TN is made smaller than the memory time m of the system, each event can no longer

    be considered a statistically-independent process. Therefore, there is a lower limit for which, in

    turn, places a lower limit on p = for a given average rate .

    The other important distribution function is obtained when Npq 1, i.e., N 1 and p and q

    are not too small. In such a case, the approximate expression for WN(K) is obtained as

    W(G)N (K) =

    12K2

    exp

    (KK)2

    2K2

    .

    (Gaussian distribution)

    (2.36)

    The mean K and variance K2 are given by (2.28) and (2.30) for the binomial distribution. When

    p 1, but the mean K = N p is much greater than one, the Poisson distribution is reduced

    to the Gaussian distribution. [The Gaussian distribution (2.36) is obtained from the binomial

    13

  • 8/3/2019 Yamamoto Noise

    17/195

    K2

    N2

    N2

    =

    Poisson

    1

    0

    0 1

    total noise

    attenuated original noise

    partition noise

    K2

    N2

    1

    0

    0

    0

    1

    N2

    = 2

    Super-Poissontotal noise

    attenuated original noise

    partition noise

    K2

    N2

    1

    0

    0 1

    N2

    = 12

    Sub-Poisson

    total noise

    attenuated original noise

    partition noise

    T

    T

    T

    Figure 2-7

    distribution (2.26) by considering WN(K) as a continuous function of K, expanding ln[WN(K)]

    around the averaged point K by a Taylor series expansion and truncating to second order.]

    2.4 Random Pulse TrainsSpectrum and Autocorrelation Function

    The Fourier transform of (2.25) is

    X(j) = F(j)Kk=1

    ak ejtk . (2.37)

    14

  • 8/3/2019 Yamamoto Noise

    18/195

    The unilateral power spectral density for such random pulse trains is defined by

    Sx() limT

    2|X(j)|2T

    = limT

    2|F(j)|2T

    K

    k,m=1

    akam exp[j(tk tm)] . (2.38)

    The summation in (2.38) over k and m can be split into the summation for k = m and k = m,

    Sx() = limT

    2|F(j)|2T

    Kk=1

    a2k +k=m

    akam exp[j(tk tm)]

    = 2a2|F(j)|2 + 4x(t)2() .

    (Carsons theorem) (2.39)

    Here, = limT

    KT is the average rate of the pulse emission and a

    2 = limT

    1K

    Kk=1

    a2k is the mean-

    square of the pulse amplitude. The mean of the noisy waveform x(t) is

    x(t) = a

    f(t)dt , (2.40)

    and a = limT

    1

    K

    Kk=1

    ak is the mean of the pulse amplitude. The term containing the delta func-

    tion represents the dc contribution to the power spectral density and is obtained by calculating

    exp[j(tk tm)] using completely random distributions of tk and tm, identifying F( = 0) with

    f(t)dt, and replacing limT

    2sin2(T/2)2T with (). For a symmetric distribution of ak about

    zero, the second term of (2.39) is zero because a is zero. However, when ak are not symmetrically

    distributed about zero, the dc term appears in the power spectral density.

    From the Wiener-Khintchine theorem, the autocorrelation function is calculated as

    x() =a2

    0|F(j)|2 cos d + 2x(t)2

    0()cos d

    = a2

    f(t)f(t + )dt + x(t)2

    , (2.41)

    15

  • 8/3/2019 Yamamoto Noise

    19/195

    where the Parseval theorem (2.12) is used to derive the second line. When = 0, one obtains

    x(t)2 x(t)2 = a2

    [f(t)]2dt

    =a2

    0

    |F(j)|2d

    (Campbells theorem of mean square) , (2.42)

    where the energy theorem (2.14) is used to derive the second line.

    2.5 Current Noise of a Vacuum Tube

    Let us consider a fluctuating current i(t) caused by random emission of electrons from the

    cathode of a vacuum tube, as shown in Fig. 2-8. Emission of electrons and subsequent travel

    of the electrons induce the surface charges on the cathode and the anode. There is a relaxation

    current flowing into a circuit to build up such a surface charge at the electrode surfaces and to

    restore the steady state of the two electrodes. This relaxation current constitutes the fluctuating

    current. In the vacuum tube case, the pulse energy (integrated relaxation current pulse) is equal

    to q (constant) and the pulse shape f(t) is determined by an electron transit velocity between the

    cathode and the anode. The pulse emission event follows the Poisson distribution with an average

    rate of emission = limT

    KT . Therefore, the average current is i(t) = q = I.

    If one assumes f(t) is an exponentially decaying function with a time constant 1, the autocor-

    relation function is calculated from (2.41) as

    i() =qI

    21e |

    |1 + I2 . (2.43)

    16

  • 8/3/2019 Yamamoto Noise

    20/195

    space charge region

    electron multiplication region

    t

    t

    t

    Poisson-point-process

    sub-Poisson-process

    reflected electron

    super-Poisson-process

    added electron

    Figure 2-8

    The unilateral power spectral density is calculated from (2.39) as

    Si() =2qI

    1 + 221+ 4I2() .

    (Schottky formula) (2.44)

    As shown in Fig. 2-9, the low-frequency power spectral density ( 11 ) is equal to the full shot

    noise 2qI and the spectral shape is Lorentzian with a cut-off frequency of 11 due to the finite

    duration of a relaxation current pulse. The dc component of the spectrum is caused by the flat

    response of the autocorrelation function I2. On the other hand, the Lorentzian noise spectrum

    stems from an exponentially decaying autocorrelation function centered at = 0.

    The physical origin for the full shot noise of the current spectrum (2.44) is the fact that the

    emission of each electron in such a thermal-limited vacuum tube is statistically independent and thus

    obeys a Poisson-point process. This is not always true. In a vacuum tube of a space-charge limited

    17

  • 8/3/2019 Yamamoto Noise

    21/195

    0 1 2 3 4 50

    1

    1 20

    2I

    q

    Si( ) 2qI

    1 -1-2-3-4

    1

    12I /q

    i()qI2( (1

    1

    Figure 2-9

    regime, there is a strong electron-electron Coulomb interaction in the space-charge region near

    the cathode. A potential minimum produced by the space charge is modulated by the fluctuating

    emission rate and provides negative feedback to regulate the effective electron emission rate, as

    shown in Fig. 2-8. In other words, each electron emission event is no longer statistically independent

    but, rather, anti-bunched. Consequently, the circuit current fluctuation is less than the full shot

    noise. This shot noise suppression due to negative feedback effect is a rather common feature in

    various electronic systems; examples will be seen in the following chapters, in which either the

    Coulomb interaction or the Pauli exclusion principle (or both) play a role in the negative feedback

    mechanism to suppress shot noise.

    On the other hand, if randomly emitted electrons are amplified by an electron multiplier, the

    initial Poissonian electron stream is transformed to the bunched electron stream, as shown in

    Fig. 2-8. Consequently, the circuit current fluctuation is more than the full shot noise. This excess

    noise is also a rather common feature in various amplifier systems, i.e., photomultipliers, avalanche

    photodiodes, and laser amplifiers.

    18

  • 8/3/2019 Yamamoto Noise

    22/195

    3. Noise figure of linear networks

    A noisy electrical network can be represented by a noise-free network with external noise

    generators. The magnitude of the external generator is expressed either by an equivalent noise

    resistance or an equivalent noise temperature. If such a network has four terminals or two (input

    and output) ports, the noise figure is often used as a figure of merit expressing the inherent noisiness

    of the circuit. The technique is also useful to express the noise properties of photonic systems such

    as lasers and optical parametric amplifiers.

    3.1 Two-terminal networks Thevenin equivalent circuit

    A noisy two-terminal network with impedance Z() = R()+jX() generates the open circuit

    voltage fluctuation v(t) as shown in Fig. 3-1(a). The two equivalent circuits based on Thevenins

    theorem[1] are shown in Fig. 3-1(b) and (c). One is the noise-free network with impedance Z()

    v(t)Z() Z() Y() i(t)

    (a) (b) (c)

    v(t)

    Figure 3-1: (a) A noisy two-terminal network. (b) Thevenin equivalent circuit with an external

    voltage source. (c) Thevenin equivalent circuit with an external current source.

    1

  • 8/3/2019 Yamamoto Noise

    23/195

    in series with a voltage generator v(t). The spectral density of v(t) is expressed by

    Sv() = 4kBRn , (3.1)

    where is the absolute temperature and Rn is the equivalent noise resistance. The other is the

    noise-free network with admittance Y() = G() +j B() in parallel with a current generator i(t).

    The spectral density of i(t) is expressed by

    Si() = 4kBGn , (3.2)

    when Gn is the equivalent noise conductance.

    If the network is linear and passive, and there is no net energy flow, i.e. the circuit is at thermal

    equilibrium condition, then Rn = R() and Gn = G() = R()/[R()2 + X()2]. The noise in

    this case is equal to Johnson-Nyquist thermal noise. However, in nonlinear active circuits, or in

    non-equilibrium condition, these equalities generally do not hold.

    When the electron temperature is different from the lattice temperature, which is the case for

    hot electron devices, it is convenient to express (3.1) and (3.2) in the alternative forms:

    Sv() = 4kBnR , (3.3)

    Si() = 4kBnG , (3.4)

    where n is the equivalent noise temperature. In circuits containing shot noise sources as primary

    noise sources, it is convenient to use the expression

    Si() = 2qI , (3.5)

    2

  • 8/3/2019 Yamamoto Noise

    24/195

    where I is the terminal current and is the shot noise suppression factor. If some smoothing

    mechanisms are dominant, for instance due to space charge effect, becomes smaller than unity.

    If there is no smoothing mechanism in the system, the power spectral density is full-shot noise

    ( = 1).

    Consider the parallel RC circuit shown in Fig. 3-2(a). The thermal noise of the register is

    represented by the parallel current source i(t) with the spectral density of Si() = 4kB/R. The

    series voltage source v(t) in the Thevenin equivalent circuit shown in Fig. 3-2(b) has the spectral

    density of

    Sv() =R2

    1 + 2(CR)2Si() . (3.6)

    The frequency dependent power spectral density (3.6) is due to the impedance of the capacitor.

    i(t)

    (a) (b)

    R C R C

    vn

    Figure 3-2: (a) A parallel RC circuit with thermal noise current source. (b) The Thevenin equivalent

    circuit with thermal noise voltage source.

    Using the Wiener-Khintchine theorem, we obtain the mean-square value of the voltage generator

    v(t)2 = 12

    0

    Sv()d = kBC

    . (3.7)

    3

  • 8/3/2019 Yamamoto Noise

    25/195

    The mean-square energy stored in the parallel RC circuit is thus equal to

    1

    2Cv(t)2 =

    1

    2kB . (3.8)

    This is an example of the equipartition theorem of statistical mechanics [2]. That is, if the system

    energy is of the form of quadratic dependence on generalized coordinate, i.e. the voltage in this case,

    the average energy of the system under thermal equilibrium condition is equal to 12kB per degree

    of freedom. Notice that the noise energy is independent of the resistance R while the magnitude

    and the bandwidth of the noise spectrum are dependent on R.

    3.2 Four-terminal networks

    A network with two pairs of terminals, input and output ports, is known as a four-terminal or

    two-port network. For a noiseless four-terminal network, the currents and voltages at the terminals

    are related to each other in terms of the impedance matrix Z or the admittance matrix Y as follows:

    V1

    V2

    =

    Z11 Z12

    Z21 Z22

    I1

    I2

    , (3.9)

    I1I2

    =

    Y11 Y12

    Y21 Y22

    V1

    V2

    . (3.10)

    The subscripts 1 and 2 refer to the input and output ports, respectively, and the sign convention is

    that currents flowing into the network are positive, as shown in Fig. 3-3. The upper case letters I

    and V indicate the Fourier transforms of the current and voltage, which are in general dependent

    on frequency.

    A noisy four-terminal network is represented by an extension of Thevenins theorem[1]. In

    Fig. 3-4(a), a series noise voltage generator appears at each port. Some degree of correlation may

    4

  • 8/3/2019 Yamamoto Noise

    26/195

    v1(t)

    i1(t) i2(t)

    v2(t)

    Figure 3-3: A noiseless four-terminal network.

    exist between these generators since the same physical mechanism may be responsible, at least in

    part, for the open circuit voltage fluctuations. The dual of Fig. 3-4(a) is shown in Fig. 3-4(b), in

    v1(t)

    i1(t)

    Z

    vn1(t)

    v2(t)

    i2(t)

    vn2(t)

    v1(t)

    i1(t)

    Y v2(t)

    i2(t)

    in1(t) in2(t)

    (a) (b)

    Figure 3-4: (a) Thevenin equivalent circuit with two external series voltage generators. (b) Thevenin

    equivalent circuit with two external parallel current generators.

    which the internal noise is represented by parallel current generators.

    The current-voltage relation of a noisy four-terminal network becomes

    V1 + Vn1

    V2 + Vn2

    =

    Z11 Z12

    Z21 Z22

    I1

    I2

    , (3.11)

    I1 + In1

    I2 + In2

    =

    Y11 Y12

    Y21 Y22

    V1

    V2

    . (3.12)

    It is often more convenient to refer both external generators to the input port. The equivalent

    circuit shown in Fig. 3-5 has a series voltage generator and a parallel current generator at the

    5

  • 8/3/2019 Yamamoto Noise

    27/195

    input, for which the current-voltage characteristic is expressed by the relation

    I1 + Ina

    I2

    =

    Y11 Y12

    Y21 Y22

    V1 + Vna

    V2

    . (3.13)

    Comparing (3.12) and (3.13), the Fourier transform of the new voltage generator vna(t) should be

    v1(t)

    i1(t)

    Y v2(t)

    i2(t)

    ina(t)

    vna(t)

    Figure 3-5: The equivalent circuit of a noisy two-port with external current and voltage generators

    in the input port.

    related to that of the current generator in2(t):

    Vna = In2Y21

    . (3.14)

    The Fourier transform of the new current generator ina(t) is

    Ina = In1 Y11Y21

    In2 . (3.15)

    The arrangement of Fig. 3-5 is particularly convenient for calculating the noise figure of the

    two-port network. However, this equivalent circuit is valid only for calculating the noise in the

    output port. It does not give the correct description of the noise in the input port. This can be

    easily seen by the fact that Ina is not equal to In1.

    6

  • 8/3/2019 Yamamoto Noise

    28/195

    3.3 Noise figure of a linear two-port

    The noise figure of a two-port is defined by

    F =total output noise power per unit bandwidth

    output noise power per unit bandwidth due to input termination

    at a specific frequency and temperature. A signal input is(t) is transferred to the output through

    an input admittance Ys and noisy two-port network, as shown in Fig. 3-6. The noise in the input

    Yina

    (t)

    vna(t)

    ins

    (t)Ys

    is

    Figure 3-6: A circuit for calculating the noise figure of a two-port network.

    admittance Ys and the noisy two-port are independent, and so the noise figure of the whole system

    can be expressed as

    F =|Ins + Ina + YsVna|2

    |Ins|2

    = 1 +Sia()

    Sis()+ |Ys|2 Sva()

    Sis()+ 2Re(ivY

    s )

    Sia() Sva()

    1/2Sis()

    , (3.16)

    where Sia(), Sva() and Sis() are the power spectra ofina(t), vna(t) and ins(t), respectively, and

    iv is the normalized cross-correlation spectral density between ina(t) and vna(t).

    The power spectral densities can be expressed as

    Sia() = 4kBGni , (3.17)

    Sva() = 4kB/Gnv , (3.18)

    7

  • 8/3/2019 Yamamoto Noise

    29/195

    Sis() = 4kBGs . (3.19)

    Here Gni and Gnv are the equivalent noise conductances, which are not necessarily actual conduc-

    tances of the two-port network. On the other hand, Gs = Re(Ys) is the actual source conductance.

    The current generator ina(t) is split into two, one part of which is uncorrelated with vna(t) and the

    other part is fully correlated with vna(t). Therefore we obtain

    Ina = Inb + YcVna , (3.20)

    where Yc is the correlation admittance of ina(t) and vna(t). Since InbVna = 0, we have

    iv InaV

    na|Ina|2 |Vna|2

    1/2 = Yc|Vna|2|Ina|2

    1/2=

    YcGniGnv

    . (3.21)

    The noise figure of (3.16) is now rewritten as

    F = 1 +GniGs

    +(Gs + Gc)

    2 + (Bs + Bc)2 (G2c + B2c )

    GnvGs, (3.22)

    where Gc and Bc are the real and imaginary parts of the correlation admittance Yc. This expression

    is easily transformed into the form

    F = F0 +(Gs Gso)2 + (Bs Bso)2

    GnvGs. (3.23)

    Here, F0 = 1 +2

    Gnv(Gso + Gc) is the minimum noise figure achieved when the source admittance

    satisfies the following matching condition:

    Gs = Gso = (GnvGni B2c )1/2 , (3.24)

    Bs = Bso = Bc . (3.25)

    8

  • 8/3/2019 Yamamoto Noise

    30/195

    The conditions (3.24) and (3.25) for the source conductance and susceptance are referred to as

    noise tuning or noise matching. The four parameters Fo, Gso, Bso and Gnv completely characterize

    the noise fluctuations of the two-port network.

    3.4 Noise figure of amplifiers in cascade

    One of the important applications of the equivalent circuit discussed in the previous section

    is the overall noise figure of the system when several amplifiers are connected in cascade as shown

    in Fig. 3-7. Suppose each amplifier in the cascade is connected to a matched load, i.e. the output

    F1, G1ina1(t)

    vna1(t)

    ins(t)Ysis(t)F2, G2 loadina2(t)

    vna2(t)

    Figure 3-7: The equivalent circuit of a cascade amplifier system.

    and input admittances of adjoining amplifiers are equal. The noise figure of the whole system in

    such a case can be written as

    F =|Isn + Ina1 + YsVna1|2

    |Ins|2+|Ina2 + Y1Vna2|2

    G1|Ins|2+|Ina3 + Y2Vna3|2

    G1G2|Ins|2

    + . (3.26)

    Here Yi and Gi are the input admittance and power gain of the i-th amplifier. If Fi is the noise

    figure of the i-th amplifier defined by (3.16), the overall noise figure is

    F = F1 + (F2 1)/G1 + (F3 1)/G1G2 + . (3.27)

    9

  • 8/3/2019 Yamamoto Noise

    31/195

    Equation (3.27) is known as Friisss formula[3]. The expression indicates, that the noise figure of

    the cascade amplifier system is essentially determined by the noise figure of the first stage if the

    power gain of the first stage is sufficiently high. It is important to use a low-noise amplifier in the

    first stage in order to realize a small overall noise figure.

    References

    [1] L. Thevenin, Comptes Rend. Acad. Scie. Paris, 97, 159 (1883).

    [2] R. Lief, Fundamentals of statistical and thermal physics (McGraw-Hill, New York, 1965).

    [3] H. T. Friiss, Proc. IRE, 32, 419 (1944).

    10

  • 8/3/2019 Yamamoto Noise

    32/195

    4. Thermal Noise and Quantum Noise

    There are two types of intrinsic noise in every physical system: thermal noise and quantum

    noise. These two types of noise cannot be eliminated even when a system is perfectly constructed

    and operated. Thermal noise is a dominant noise source at high temperatures and/or low frequen-

    cies, while quantum noise is dominant at low temperatures and/or high frequencies. A conductor is

    the simplest physical system which produces these two types of intrinsic noise. The intrinsic noise

    of macroscopic and mesoscopic conductors will be discussed in this chapter.

    A conductor in thermal equilibrium with its surroundings shows, at its terminals, open-circuit

    voltage (or short-circuit current) fluctuations, as shown in Fig. 4-1. Thermal equilibrium noise

    R

    R

    v

    i

    Sv( ) = 4k B R Si( ) =4kB

    R

    Figure 4-1:

    was first observed by M. B. Johnson in 1927. He discovered that the noise power spectral density

    is independent of both the material the conductor is made of and the measurement frequency,

    and is determined only by the temperature and electrical resistance. The open-circuit voltage

    noise spectral density is Sv() = 4kBR and the short-circuit current noise spectral density is

    Si() = 4kB/R. This noise is referred to as thermal noise and is the most fundamental and

    1

  • 8/3/2019 Yamamoto Noise

    33/195

    important noise in electronic devices.

    The physical origin of thermal noise in a macroscopic conductor is a Mrandom-walk of

    thermally-fluctuated electrons. An electron undergoes a Brownian motion via collisions with the

    lattices in a conductor. The statistical properties of such a Brownian particle were studied first

    by A. Einstein twenty years before Johnsons observation of thermal noise. The electrons in a

    conductor are thermally energetic via collisions with the lattice and travel randomly. The electron

    velocity fluctuation is a statistically-stationary process. However, the mean-square displacement

    of an electron increases in proportion to the observation time. The electron position fluctuation is

    a statistically-nonstationary process. Such a microscopic approach can indeed explain Johnsons

    observation.

    Nyquist employed a completely different approach to the problem. He introduced the concept

    of a mode for the system by using a transmission line cavity terminated by two conductors. He

    then applied the equipartition theorem of thermodynamics to these transmission line modes. In

    this way he could explain Johnsons observation without going into the details of a microscopic

    electron transport process. Nyquists approach is very general and is easily extended to include

    quantum noise, which is important in a high-frequency and low-temperature case.

    In order to introduce quantum noise to a conductor, one must assume the quantization pos-

    tulate. The current and voltage of a LC circuit are a pair of conjugate observables just like the

    position and momentum of a particle or a mechanical harmonic oscillator. One cannot determine

    these two conjugate observables simultaneously with arbitrary accuracy due to the Heisenberg

    uncertainty principle. These quantum uncertainties in the current and voltage have their micro-

    2

  • 8/3/2019 Yamamoto Noise

    34/195

    scopic origins in that the momentum (corresponding to the short circuit current) and position

    (corresponding to the open circuit voltage) of an electron in a conductor cannot be determined

    simultaneously.

    A mesoscopic conductor does not impose any scattering on electrons and an electron travels as

    a coherent wave in such a small-scale device. Nevertheless, there is still a finite electrical resistance

    (quantum unit of resistance RQ =h2e2 ) due to the Pauli exclusion principle. It is truly surprising to

    find that the formula for thermal noise and quantum noise of a macroscopic conductor with many

    scatterings still hold for such a mesoscopic conductor.

    4.1 Microscopic Theory of Thermal Noise Einsteins Approach

    4.1-1 Mean Free-Time and Mobility

    u(t)

    collision

    f1f2f3 E E E E E E E

    u(t) : average driftvelocity

    V

    collision with the lattice

    t

    Figure 4-2:

    Consider a one-dimensional conductor under an applied voltage (Fig. 4-2). We assume that an

    electron is accelerated by the uniform electric field between collisions with the lattice and that an

    3

  • 8/3/2019 Yamamoto Noise

    35/195

    electron velocity returns to zero at every collision event. This is not true in a real collision process

    in a conductor, but the conclusion one obtains using this assumption is essentially same as that

    obtained by more realistic collision models. Since the electron drift velocity in this model is given

    by u(t) = t = qEm

    t during the time between collisions f, the displacement between two collisions

    is

    x(f) =

    22f =

    qE

    2m2f . (4.1)

    After K collisions with the lattice, the total displacement is qE2m 2f K. Therefore, the mean drift

    velocity u is given by

    u total displacementtotal time

    =

    qE2mK

    2f

    Kf=

    q2f2mf

    E . (4.2)

    Here, f and 2f are the mean free time and mean-square free time. The mobility is defined by

    u = E. Thus, one obtains

    =q2f

    2mf. (4.3)

    As shown in Fig 4-2, f is randomly distributed around its mean value f. The probability

    pi(m, ) that the ith electron experiences exactly m collisions in a time interval [0, ] obeys a Poisson

    distribution if the probability of electron collision with the lattice is independent of the electron

    drift velocity which is a reasonable assumption for a weak field; thus it is a statistically independent

    process. Strictly speaking, the electron collision with a lattice obeys a binomial distribution with

    a very small collision probability starting from a very large number of (potential) collision events,

    which is well approximated by a Poisson distribution. Thus, one obtains

    pi(m, ) =(i)m

    m!ei , (4.4)

    4

  • 8/3/2019 Yamamoto Noise

    36/195

    where i is the mean rate for collision per second. The probability for a free time f lying between

    fi and fi + dfi is equal to the joint probability of zero collisions in a time interval [0, fi] and

    one collision in a time interval [fi, fi + dfi]. Thus,

    q(fi)dfi = pi(0, fi) pi(1, dfi)

    = i eifi dfi , (4.5)

    where eidfi 1 is used and i eifi is considered a probability density function by which we

    can calculate the mean free time and mean-square free time,

    fi = i

    0

    fi eifi dfi = 1

    i, (4.6)

    2fi = i

    0

    2fi eifi dfi =

    2

    2i= 2fi

    2 . (4.7)

    As seen from (4.6) and (4.7), fi does not obey a Poisson distribution, but, rather, a geometrical

    distribution. Figure 4-3 compares pi(m, ) and qi(fi). If N electrons behave in a similar but

    0

    0.5

    1

    0 10

    0.5

    1

    0 1 2 3 4 5

    Poisson distribution

    m

    Geometical distribution

    i fi i

    2 i p i m,qi fi / i

    Figure 4-3:

    independent way, the Maddition theorem of a Poisson process is applied and the total probability

    of collision p(m, ) still obeys a Poisson distribution with the mean rate =Ni=1

    i =Nf

    . The

    (collective) mean free time f is defined by f =N

    .

    5

  • 8/3/2019 Yamamoto Noise

    37/195

    Using (4.7) in (4.3), the mobility is uniquely related to the mean free time,

    =qfm

    . (4.8)

    According to the collision model, any departure of the drift velocity u(t) = u(t)u from its mean

    value decays with a time constant f =mq

    .

    4.1-2 Velocity and Current Fluctuations

    In a thermal equilibrium condition (zero applied voltage), the mean drift velocity of an electron

    is zero, u = 0. However, an electron acquires a non-zero momentum when it collides with a lattice

    which decays toward the mean value of u = 0 with a time constant f. If its momenta before and

    after a collision are p1 and p2, respectively, the collision imparts a momentum change (p2 p1) in

    an infinitesimal time. The equation of motion is then

    du(t)

    dt= u(t)

    f+

    (p2 p1)(t)m

    . (4.9)

    The drift velocity u(t) is kicked randomly by the second part of the RHS of (4.9) and is simulta-

    neously damped by the first part of the RHS of (4.9). The Fourier analysis of (4.9) requires the

    introduction of a gated function uT(t) = u(t) for T2 t T2 and 0 for otherwise because u(t) is

    a statistically-stationary process. The Fourier transform of this gated function leads to

    UT(j) =

    q

    (p2 p1)1 +jf

    . (4.10)

    From the Carson theorem, the power spectral density of the velocity fluctuation is given by

    Su() = 2(p2 p1)2 (/q)2

    1 + 22f, (4.11)

    6

  • 8/3/2019 Yamamoto Noise

    38/195

    where = limT

    K

    T=

    1

    fis the mean rate of collisions, K is the number of collision events, and

    (p2 p1)2 = limT

    1

    K

    K

    k=1

    (p2k p1k)2 is the mean square of the momentum change per collision.

    Since momenta before and after a collision are statistically independent,

    (p1 p2)2 = 2p2

    = 2 limT

    1

    K

    Kk=1

    p2k

    . (4.12)

    For a thermal equilibrium condition, the mean square of the electron momentum is determined by

    the equipartition theorem:

    p2

    2m

    =1

    2

    kB (one

    dimensional case) . (4.13)

    From (4.11)(4.13), one obtains

    Su() =4kB/q

    1 + 22f, (4.14)

    where (4.8) is used.

    From the Wiener-Khintchine theorem, one can calculate the autocorrelation function

    u() =1

    2

    0

    Su()cos d

    =kB

    qfexp

    ||

    f

    . (4.15)

    Therefore, the mean square value of u(t) is given by u(0) =kBqf

    . Figures 4-4(a) and (b) show

    Su() and u().

    The velocity fluctuation u(t) of the electron produces a short-circuited current fluctuation i(t) =

    qu(t)/L in an external circuit, where L is the length of the conductor. Since the current fluctuation

    of each electron is additive, the total current fluctuation power spectral density is given by

    Si() = Su()q2

    L2ALn , (4.16)

    7

  • 8/3/2019 Yamamoto Noise

    39/195

    where A is a cross-sectional area of the conductor and n is the electron density. If one uses (4.14)

    and the expression for electrical resistance, R = LA

    = LnqA

    , in (4.16), one obtains

    Si() =4kB/R

    1 + 22f. (4.17)

    Therefore, the short-circuited current fluctuation power spectral density at low frequency is 4kB/R,

    which is exactly what Johnson observed experimentally.

    1

    0.5

    0 1-1101

    1

    Su( ) /4kB

    q

    f

    u )q f

    k

    f

    10-1

    10-110

    -2

    10-2

    Figure 4-4:

    4.1-3 Position and Charge Fluctuations

    The integral of a statistically-stationary process which shows a white power spectral density

    is called a Wiener-Levy process. The surface charge induced by the fluctuating short-circuited

    current across a conductor is

    q(t) =

    t0

    i(t)dt . (4.18)

    Consider again the one-dimensional conductor shown in Fig. 4-2. An electron transit over a free

    path f between collisions gives rise to a surface charge on the electrode equal to qfL

    . If there

    are m = t independent events in a time duration t , then the mean-square value of the surface

    8

  • 8/3/2019 Yamamoto Noise

    40/195

    charge q(t) is calculated as

    q(t)2 = q2(t)2fL2

    =2kB

    Rt , (4.19)

    where = nAL/f is the mean rate of collisions, 2f = v

    2T

    2f = 2kB f/q is the mean-square

    free path, and v2T = kB/m is the mean-square thermal velocity; therefore, the charge fluctuation

    is not stationary, but, rather, diffuses with time. The diffusion coefficient, defined by q(t)2 = 2Dt,

    is given by

    Dq =kB

    R=

    1

    4Si( 0) . (4.20)

    Since q(t) is a cumulative process, one can write

    q(t + ) = q(t) + q(t, ) , (4.21)

    where q(t, ) is the charge fluctuation added between times t and (t + ). Since the correlation

    time of the additive fluctuations is very short (f ), the two processes in the RHS of (4.21) are

    uncorrelated; therefore, the covariance function is equal to the mean square value:

    q(t + )q(t) = q(t)2 + q(t)q(t, ) q(t)2 . (4.22)

    The autocorrelation function must be redefined in such a manner that the observation time remains

    finite rather than infinite and the ordinate is set to zero outside the observation time interval. Thus,

    the autocorrelation function is a function of the time delay and the observation time T:

    q(, T) =1

    T

    T||0

    q(t + )q(t)dt

    = 1T

    T||0

    2kBR

    t dt

    =kB

    RT

    1 ||

    T

    2. (4.23)

    9

  • 8/3/2019 Yamamoto Noise

    41/195

    The power spectral density is obtained by the Wiener-Khintchine theorem as:

    Sq(, T) = 4 T

    0q(, T)cos d

    =8kB

    2R

    1 sin(T)

    T

    . (4.24)

    When T 1, the power spectral density is proportional to 2, which is characteristic of a

    Wiener-Levy process. Figures 4-5(a)(c) show the real-time function, autocorrelation function,

    and power spectral density of the surface charge q(t). The integral of an electron velocity u(t)

    t

    1

    1 10

    q(t)

    0

    random walk diffusion

    =

    0.1

    S , T)q

    0-1 1

    1

    0.5

    , T)q.

    q(t)2 2kB

    Rt

    3R

    4kB T2

    -2

    -2

    -1

    R

    k T

    Figure 4-5:

    is an electron displacement x(t), which is also a Wiener-Levy process. By performing the same

    10

  • 8/3/2019 Yamamoto Noise

    42/195

    calculation, one obtains the (position) diffusion coefficient,

    Dx =kB

    q=

    1

    4Su( 0) .

    (Einsteins relation) (4.25)

    4.2 Macroscopic Theory of Thermal Noise Nyquists Approach

    Nyquists treatment of thermal noise appeared very soon after Johnsons observation and the

    result is often referred to as Nyquists Law. Nyquist initially considered two electrical conductors,

    R1 and R2, connected in parallel, as shown in Fig. 4-6. The open-circuit voltage noise source v1

    associated with conductor R1 produces a current fluctuation in the circuit, leading to an absorbed

    power R2v21/(R1 + R2)2 by conductor R2. A similar flow of absorbed power, R1v22/(R1 + R2)

    2,

    exists from R2 to R1. Since the two conductors are at the same temperature, the power flow in

    each direction must be exactly the same and cancel each other out; otherwise, the second law of

    thermodynamics would be violated. The second law of thermodynamics states that it is impossible

    to take heat from one reservoir to another reservoir at an equilibrium temperature. In order to

    satisfy this exact cancellation of power flow from R1 to R2, and vice versa, the open circuit voltage

    noise v2i should be proportional to the electrical resistance Ri. This exact cancellation of power flow

    must hold not only for the total power, but also for the power exchanged in any frequency band;

    otherwise, the second law of thermodynamics would be violated simply by inserting a frequency

    filter between the resistors. In other words, the power spectrum Sv() of the voltage fluctuations

    should be independent of the detailed structure and material of the conductor and should be a

    universal function ofR, , and (angular) frequency . Specifically, if the conductors R1 and R2 are

    11

  • 8/3/2019 Yamamoto Noise

    43/195

    at different temperatures, 1 and 2, the net heat flow should be proportional to the temperature

    difference 12 and thus the power spectrum Sv() is proportional to the temperature . Nyquist

    R1conductor 1

    conductor 2 2vv1

    R2

    R2

    R1 + R22 v1

    2

    )(

    R

    R1 + R22 v

    2

    )(1

    2

    Figure 4-6:

    then considered a long, lossless transmission line terminated at either end by conductors R1 and

    R2 (Fig. 4-7). This lossless LC transmission line has a characteristic impedance Z0 =LC

    which

    is equal to the end terminal conductance R1 = R2 = R and the wave velocity v =1LC

    , where

    L and C are the inductance and capacitance of the transmission line per unit length. The power

    delivered to the transmission line from R1 or R2 in a frequency interval d/2 is given by

    dP =1

    4RSv()

    d

    2. (4.26)

    A time duration in which this noise power travels in the transmission line is t = v

    , where is the

    length of the transmission line. Hence, the total stored energy in the transmission line in the same

    frequency band is

    dE = dP t 2 = 2Rv

    Sv()d

    2. (4.27)

    Suppose the transmission line is suddenly short-circuited by closing the two switches across R1

    and R2 and, therefore, the energy density (4.27) on the transmission line is trapped as standing

    waves. The resonant frequencies of these standing waves are N =v2 N, where N is a positive

    12

  • 8/3/2019 Yamamoto Noise

    44/195

    1=

    v

    2l

    2=

    v

    2l2~

    3

    =v

    2l3~

    N=

    v

    2lN~

    R 1

    2v

    R2

    l

    v1

    Figure 4-7:

    integer. The number of standing wave modes in the frequency interval d2 is given by

    m =

    d

    2

    v

    2

    =

    d v

    . (4.28)

    As the transmission line length goes to infinity, the number of degrees of freedom (DOF) of the

    system (given by the number of the standing wave modes) also goes to infinity. Therefore, it is

    permissible to invoke the equipartition theorem to determine the total energy in the transmission

    line at thermal equilibrium. The average energy per mode (with two DOFs) is kB and thus the

    stored energy in the frequency band d2 is given by

    dE = mkB = dv

    kB . (4.29)

    13

  • 8/3/2019 Yamamoto Noise

    45/195

    From 4.27) and 4.29), the power spectrum is obtained as

    Sv() = 4kBR . (4.30)

    As expected from a thermodynamic argument, Sv() is proportional to R and . If an angular

    frequency becomes greater than kB/h, the average thermal energy per DOF is calculated as

    h nth =h

    exphkB

    1

    , (4.31)

    where the photon number distribution function Pn obeys the geometrical distribution

    Pn =nnth

    (1 + nth)n+1= exp

    nh

    kB

    1 exp

    h

    kB

    , (4.32)

    and thus the mean thermal photon number is

    nth n=0

    nPn =1

    exphkB

    1

    . (4.33)

    Therefore, the total energy in the frequency band d2 is

    dE = mhnth =d

    v h

    exphkB

    1

    . (4.34)

    Comparing this result with (4.27), the power spectral density of the voltage noise is found to be

    Sv() =4hR

    exphkB

    1

    . (4.35)

    In (4.35), the quantized energy h of a photon is taken into consideration. However, the full

    quantum mechanical analysis of the problem suggests that it is still insufficient since the zero-point

    fluctuation is not included.

    14

  • 8/3/2019 Yamamoto Noise

    46/195

    4.3 Quantum Noise

    Consider a lossless LC circuit in order to derive the zero-point fluctuation. The resonant

    frequency of an LC circuit is given by 0 =1LC

    . One now models each transmission line cavity

    mode in Fig. 4-7 by such a lumped LC resonant circuit. If the current flowing in the inductance L

    is denoted by i(t) and the voltage across the capacitance C is denoted by v(t), the Kirchhoff laws

    are represented by

    Cdv(t)

    dt= i(t) , (4.36)

    L

    di(t)

    dt = v(t) . (4.37)

    These equations can be expressed in terms of the normalized voltage, q(t) Cv(t), and the

    normalized current, p(t) Li(t):dq(t)

    dt=

    1

    Lp(t) , (4.38)

    dp(t)

    dt= 1

    Cq(t) . (4.39)

    The total energy stored in the lossless LC circuit is given by

    H =1

    2Li2 +

    1

    2Cv2 =

    p2

    2L+

    q2

    2C. (4.40)

    Suppose one interprets the above expression in terms of an analogy with a mechanical harmonic

    oscillator with a mass m and oscillation frequency 0 =1LC

    . One can then identify the inductance

    L corresponding to the mass m, the capacitance C corresponding to the inverse of the spring

    constant 1k

    = 1m2

    0

    , and p and q corresponding to the momentum and position, respectively. If one

    considers 4.40) as a Hamiltonian function of the system, the classical Hamilton equations for the

    15

  • 8/3/2019 Yamamoto Noise

    47/195

    position q and momentum p are written as

    dq

    dt H

    p=

    1

    Lp , (4.41)

    dp

    dt H

    q= 1

    Cq , (4.42)

    which are identical in form to the Kirchhoff equations (4.37) and (4.38) for the voltage q and

    current p. Hence, one can conclude that the voltage and current in a lossless LC circuit

    are a pair of conjugate observables, just like the position and momentum of a mechanical har-

    monic oscillator. Quantum mechanically, conjugate observables q and p must satisfy the following

    commutation relation,

    [q, p] qp pq = ih , (4.43)

    where q and p are no longer complex numbers (c-numbers); rather, they are quantum mechanical

    operators (q-numbers). Using the Schwartz inequality, one can derive the following Heisenberg

    uncertainty principle for the product of the variances of q and p:

    q2

    p2

    h2

    4 . (4.44)

    The (non-Hermitian) creation (a) and annihilation (a) operators are defined by:

    a =1

    2h0L(0Lq + ip) , (4.45)

    a =1

    2h0L(0Lq ip) . (4.46)

    The creation operator a is a Hermitian conjugate of the annihilation operator a. The Hamiltonian

    function (4.40) and the commutation relation (4.43) are rewritten as

    H = h0

    aa +

    1

    2

    , (4.47)

    16

  • 8/3/2019 Yamamoto Noise

    48/195

    [a, a] = 1 . (4.48)

    If the LC circuit is not excited at all, the quantum mechanical state of such an unexcited LC circuit

    is referred to as the ground state, or the vacuum state. The vacuum state is mathematically

    defined by

    a|0 = 0 . (4.49)

    One cannot annihilate a photon from the vacuum state because there is no photon in the vacuum

    state, which is the meaning of this expression. On the other hand, the operation of the creation

    operator on the vacuum state creates a single-photon state:

    a|0 = 1 . (4.50)

    The mean and mean-square of the voltage and current of the vacuum state are calculated by

    taking the ensemble average of the respective operators:

    q 0|q|0 = 0|

    h0C

    2(a + a)|0 = 0 , (4.51)

    q2 0|q2|0 = 0| h0C2

    (a2 + a2 + aa + aa)|0 = h02

    C , (4.52)

    p 0|p|0 = 0|i

    h0L

    2(a a)|0 = 0 , (4.53)

    p2 0|p2|0 = 0| h0L2

    (a2 a2 + aa + aa)|0 = h02

    L . (4.54)

    Here, the orthogonality relations between photon number eigenstates such as 0|1 = 0|2 = 0 are

    used. The variances of the voltage and current satisfy the minimum uncertainty product,

    q2 q2 q2 = h02 C

    p2 p2 p2 = h02 L q2 p2 =

    h2

    4. (4.55)

    17

  • 8/3/2019 Yamamoto Noise

    49/195

    Notice that both voltage and current have intrinsic fluctuations (variances) even at absolute zero

    temperatures. The expectation value of the zero-point energy is given by

    H 0|H|0 = 0|h0

    aa +1

    2

    |0 = h0

    2. (4.56)

    If one adds this zero-point fluctuation contribution to (4.35), one obtains the full quantum

    mechanical expression for an open-circuit voltage fluctuation spectral density:

    Sv() = 2hR coth

    h

    2kB

    . (4.57)

    As shown in Fig. 4-8, Sv() is reduced to the thermal noise value (4kBR) in the high-temperature

    limit (kB h) and is reduced to the quantum noise value (2hR) in the low-temperature limit

    (kB h). Equation (4.57), including the quantum mechanical zero-point fluctuation, is referred

    to as the generalized Nyquist noise.

    1

    10

    1 10

    102

    10-1

    10-1

    10-2 102

    Thermal Noise limit

    Sv() / 4k BR

    Quantum Noise Limit

    h

    2kB

    Figure 4-8:

    18

  • 8/3/2019 Yamamoto Noise

    50/195

    4.4 Thermal and Quantum Noise of a Mesoscopic System

    If the size of a conducting wire becomes smaller and smaller, an electron can eventually transit

    from one electrode to the other without a collision with the lattice. Such a small-scale conductor is

    called a mesoscopic conductor. A collision-free electron transport in a mesoscopic conductor is

    termed a ballistic transport, while a conventional electron transport accompanying many collisions

    with the lattice in a macroscopic conductor is termed a dissipative transport. Since an electron

    propagates as a coherent wave in such a mesoscopic conductor, various coherent electron wave

    interferometric devices can be constructed. Even though such an electron wave is not scattered by

    the lattice, it is known that the terminal current of such a mesoscopic conductor still features noise.

    Consider a mesoscopic conductor with two terminal electrodes connected to an external constant

    V

    electrode electrodemesoscopic conductor

    electronenergy

    coherent electron wave

    kN =2

    LN

    1

    2

    1 2 qV

    Figure 4-9:

    voltage source, as shown in Fig. 4-9. The electron distributions in the two terminal electrodes

    obey the thermal equilibrium Fermi-Dirac distributions. The Fermi-level difference is determined

    by the external voltage. An electron wavepacket with a certain energy is emitted from electrode 1

    19

  • 8/3/2019 Yamamoto Noise

    51/195

    and absorbed by electrode 2, which contributes to an external current. As shown in Fig. 4-9, the

    longitudinal wavenumber of such a coherent electron wave is quantized by the boundary condition:

    kN =2

    LN (N = positive integer) . (4.58)

    The mode density per energy is given by

    dN

    dE=

    dN

    dkN

    dkNdE

    =L

    2hvN, (4.59)

    where vN =ddkN

    is the electron group velocity. The total number of independent modes in the

    energy range 1 2 is dNdE (1 2). Each mode can accommodate a single electron due to the

    Pauli exclusion principle and has a transit time tN =LvN

    between the two electrodes. Therefore,

    the average current is

    I = qdN

    dE(1 2) 1

    tN=

    q2

    hV , (4.60)

    where 1 2 = qV (V = external bias voltage) is used. A mesoscopic conductor has a finite

    conductance GQ =q2

    h(quantum unit of conductance) that is not due to scattering with the lattice

    (dissipation), but due to the Pauli exclusion principle.

    An alternative method to derive the quantum unit of conductance GQ is to count the number

    of independent temporal modes (rather than Fourier modes). As shown in Fig. 4-10, the

    inverse Fourier transform of a bandpass-filtered spectrum centered at a frequency 0 and bandwidth

    (FWHM) = 12h

    is a Nyquist function fN =sin(t)(t) . A sequence of Nyquist functions

    displaced by = 1 reproduces any bandwidth-limited function, and thus one can consider the

    arrival rate of independent temporal modes as 1

    = . If one assigns a single electron per

    20

  • 8/3/2019 Yamamoto Noise

    52/195

    independent temporal mode, the current is

    I =q

    =

    q2

    hV . (4.61)

    The same result is obtained as in 4.60). If a mesoscopic conductor supports only one transverse

    mode, two spin-degenerate states can still be accommodated for the single transverse mode and

    thus the total conductance is 2q2

    hinstead of q

    2

    h. If a system is at an absolute zero temperature,

    Spectrum

    0 E=

    1

    Nyquist

    function

    12 t

    Figure 4-10:

    as shown in Fig. 4-11(a), the state above the Fermi energy is completely empty and the state

    below the Fermi energy is completely occupied. The electron wavepackets (temporal modes) with

    center frequencies lying between the two Fermi energies are emitted from electrode 1 and travel

    toward electrode 2. The electron wavepackets with energies outside of this region do not contribute

    to a current. This electron transport is a deterministic process and thus there is no dc current

    noise in the external circuit. However, if one measures the current noise at a finite frequency, the

    noise is produced by the beating between the electron wavepacket emitted from one electrode in

    the energy below the Fermi energy 1 and the empty electron wavepacket (vacuum state) emitted

    from the same electrode in the energy above the Fermi energy 1. By increasing the measurement

    21

  • 8/3/2019 Yamamoto Noise

    53/195

    frequency, additional occupied and empty electron wavepackets contribute to the beat noise and

    the noise power thus increases linearly in proportion to . This is the origin of the quantum noise

    which is dominant in a low-temperature/high-frequency region, as shown in Fig. 4-12.

    If a system is at a finite temperature, as shown in Fig. 4-11(b), the electron occupancy near

    the Fermi energy becomes probabilistic. The electron wavepackets with energies close to 1 and

    2 may or may not be emitted from electrode 1 and absorbed by electrode 2. This stochasticity

    introduces another current noise in the external circuit. By increasing the temperature, the partial

    occupancy of electron states increases and the noise power thus increases linearly in proportion to

    . This is the origin of the thermal noise which is dominant in a high-temperature/low-frequency

    region, as shown in Fig. 4-12.

    beating

    empty packets

    occupiedpackets beating

    (a) = 0 (Quantum Noise) (b) = 0 (Thermal Noise)

    statistical occupancy

    2

    1

    2

    1

    Figure 4-11:

    22

  • 8/3/2019 Yamamoto Noise

    54/195

    1

    10

    1 10

    10

    101010 10

    Thermal Noise limit

    Quantum Noise

    h

    2kB

    S ( )RQ

    4kB

    Figure 4-12:

    References

    1. Mobility and diffusion constant of electrons in conductors:

    W. Shockley, Electrons and Holes in Semiconductors, Van Nostrand, New York (1963).

    2. Langevin equation and fluctuation-dissipation theory of a Brownian particle:

    M. Sargent III, M. O. Scully and W. E. Lamb, Jr., Laser Physics (Addison-Wesley,

    Reading, Mass. 1974).

    3. Thermal noise:

    M. B. Johnson, Nature 119, 50 (1927); Phys. Rev. 32, 97 (1928).

    H. Nyquist, Phys. Rev. 32, 110 (1928).

    4. Quantum zero-point fluctuations of a loss less LC circuit:

    23

  • 8/3/2019 Yamamoto Noise

    55/195

    W. H. Louisell, Radiation and Noise in Quantum Electronics, McGraw-Hill Book Co.,

    New York (1964); Quantum Statistical Properties of Radiation, John Wiley & Sons, New

    York (1973).

    5. Generalized Nyquist formula:

    H. B. Caller and T. A. Welton, Phys. Rev. 83, 34 (1951).

    24

  • 8/3/2019 Yamamoto Noise

    56/195

    Contents

    5. Inherent Noise of pn Junction Diodes 2

    5.1 pn Junction Diodes Under Constant Voltage Operation . . . . . . . . . . . . . . . . 4

    5.1.1 Current-Voltage and Capacitance-Voltage Characteristics of a pn Junction . . 4

    5.1.2 Thermal Diffusion Noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

    5.1.3 Generation-Recombination Noise . . . . . . . . . . . . . . . . . . . . . . . . . 13

    5.1.4 Total Current Noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

    5.1.5 Short Diode . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

    5.2 pn Junction Diodes Under Constant Current Operation . . . . . . . . . . . . . . . . 21

    5.2.1 Effect of a Finite Source Resistance . . . . . . . . . . . . . . . . . . . . . . . 21

    5.2.2 A pn Junction Diode With Negligible Backward Thermionic Emission . . . . 23

    1

  • 8/3/2019 Yamamoto Noise

    57/195

    5. Inherent Noise of pn Junction Diodes

    Shockleys 1949 paper heralded a new era in the history of semiconductor device physics.

    Basic physical processes of a junction diode and a junction transistor were presented in this 1949

    Bell Syst. Tech. J. paper. In the same issue, the first report appeared on the noise of a point

    contact transistor. The observed noise figure was 50-70 dB above the intrinsic noise limit! It took

    almost 30 years to suppress this excess noise (mainly due to 1/f noise and surface recombination

    noise) and to obtain a noise figure very close to the theoretical limit (= 0 dB). This intrinsic noise

    of a transistor is determined by the thermal noise in the bulk resistive region and the shot noise in

    the junction region. In this chapter we will study the inherent noise of pn junction diodes, which

    set a fundamental limit on the noise performance of various semiconductor devices.

    There are two distinct bias conditions for a junction diode: constant voltage operation and

    constant current operation. The former is obtained when the junction (differential) resistance Rd

    is much larger than the source resistance RS, and the latter is obtained in the opposite limit.

    There are two types of junction diodes: a macroscopic junction and a mesoscopic junction. An

    electrostatic energy for a single electron q2/2C (where C is a junction capacitance) is much smaller

    than the thermal energy kB in a macroscopic junction and the opposite is true for a mesoscopic

    junction. A junction diode features markedly different noise characteristics in such different bias

    conditions and operational regimes.

    Consider a pn junction diode biased by a constant voltage source with a source resistance RS,

    as shown in Fig. 5-1. If the source resistance RS is much smaller than a diode differential resistance

    2

  • 8/3/2019 Yamamoto Noise

    58/195

    defined by

    Rd

    dI

    dV

    1

    , (5.1)

    where I and V are the junction current and the junction voltage, then the junction voltage V is

    p

    n

    Rs

    v

    RdCdep CdifRsis i

    Figure 5-1:

    pinned by the source. There is no fluctuation in the junction voltage V due to the fast relaxation

    time (CRS) of an external circuit, but there is a fluctuation in the junction current I. This bias

    condition is referred to as constant voltage operation. Standard theoretical studies on the noise

    characteristics of a pn junction diode have considered this mode of operation; therefore, our analysis

    also starts with this bias condition (Section 5.1).

    On the other hand, some experimental studies on the noise of a pn junction diode have consid-

    ered the opposite bias condition; that is, the source resistance RS is much larger than the differential

    resistance Rd of the diode. In such a case there is no fluctuation in the junction current I due to the

    slow relaxation time (CRS) of an external circuit, but there is a fluctuation in the junction voltage

    V. This bias condition is referred to as constant current operation. This mode of operation will

    be analyzed in Section 5.2.

    There are two types of pn junctions which feature drastically different noise characteristics:

    macroscopic junctions and mesoscopic junctions. When a single-electron charging energy (q2/2C)

    3

  • 8/3/2019 Yamamoto Noise

    59/195

    (where C is the junction capacitance) is much smaller than the thermal characteristic energy

    kB, the behavior of each individual electron does not affect the junction dynamics. This is a

    macroscopic junction limit. On the other hand, when (q2/2C) is much greater than kB, a so-called

    (single-electron) Coulomb blockade effect emerges and a previous single-electron event determines

    the junction dynamics at a later time. This is a mesoscopic junction limit.

    5.1 pn Junction Diodes Under Constant Voltage Operation

    5.1.1 Current-Voltage and Capacitance-Voltage Characteristics of a pn Junction

    The noise characteristics of a p+-N heterojunction with a heavily p-doped narrow bandgap

    material and lightly n-doped wide bandgap material will be studied, rather than a conventional

    p-n homojunction, because this specific junction structure is used in various semiconductor devices

    (i.e., a double-heterostructure semiconductor laser and a heterojunction bipolar transistor). The

    extension of the following analysis to a pn homojunction is straightforward.

    The band diagram of a p+

    -N heterojunction diode at a zero bias condition (V = 0) and a

    forward bias condition (V > 0) are shown in Fig. 5-2. The built-in potential VD is divided into the

    potentials in the p+- and N-layers:

    VDp =VDK

    , (5.2)

    VDn = VD

    1 1

    K

    , (5.3)

    where

    K = 1 +1(N

    A1 N+D1)2(N

    +D2 NA2)

    . (5.4)

    4

  • 8/3/2019 Yamamoto Noise

    60/195

    N-layer

    Ecq

    V

    electron gas

    +

    +

    +

    +

    +

    +

    x

    +

    +

    x

    electron gas

    +

    hole gas hole gas

    x = -d x = 0

    (v = 0)

    x = -d x = 0

    (v > 0)

    EVq

    VDn

    VDp

    VDn - V2

    VDp - V1

    F

    p+

    - layer

    Fp

    Fn

    Figure 5-2:

    Here, 1, NA1, and ND1 are the dielectric constant, acceptor concentration, and donar concentration

    of the p+-layer, and 2, NA2, and ND2 are those of the N-layer. If a p+-N diode satisfies 1 > 2

    and NA1 N+D1 N+D2 NA2, then K 1. Consequently, the built-in potential VD is mainly

    supported in the N-layer, i.e., VDn VD and VDp 0. The transmitted electron flux from theN-layer to the p+-layer across the potential barrier height VDn should be equal to the transmitted

    electron flux from the p+-layer to the N-layer across the potential barrier Ec/q because there is

    no net current at V = 0.

    When a forward bias (V > 0) is applied, only the potential barrier seen by the electrons in the

    N-layer decreases to VDn V2 VD V, where the applied voltage supported in the N-layer is

    V2 = V

    1 1K

    V. The electron (minority carrier) density at the edge of the depletion layer

    5

  • 8/3/2019 Yamamoto Noise

    61/195

    (x = 0) in the p+-layer is given by

    np = XnN0 expVD V

    VT = np0 expV

    VT , (5.5)where VT =

    kBq is the thermal voltage and

    np0 = XnN0 exp

    VD

    VT

    , (5.6)

    is the thermal equilibrium electron density in the p+-layer. X is the transmission coefficient of

    an electron at the heterojunction interface and nN0 is the electron (majority carrier) density at

    the edge of the depletion layer (t =

    d) in the N-layer which is equal to the thermal equilibrium

    electron density in the N-layer.

    The excess electron density 5.5) at x = 0 diffuses towards x = W, where a p-side metal contact

    is located. The distribution of the excess electron density n(x, t) obeys

    tn(x, t) = n(x, t) np0

    n 1

    q

    xin(x, t) , (5.7)

    where n is the electron (minority-carrier) lifetime and, since there is no electric field in the neutral

    p+-layer, the current in(x, t) is carried only by a diffusion component

    in(x, t) = qDn x

    n(x, t) . (5.8)

    Here, Dn is the electron diffusion constant. Solving (5.7) and (5.8) with the boundary conditions,

    np =

    np0 expVVT

    at x = 0

    np0 at x

    Ln

    , (5.9)

    the steady-state solution for n(x) is now given by

    np(x) = np0 + (np np0)ex/Ln , (5.10)

    6

  • 8/3/2019 Yamamoto Noise

    62/195

    where Ln =

    Dnn is the electron diffusion length. The junction current density is determined by

    the diffusion current (5.8) at x = 0:

    i in(x = 0) = qDnLn

    (np np0) = qDnnp0Ln

    eVVT 1

    . (5.11)

    The total current I = Ai vs. the junction voltage V is plotted in Fig. 5-3, where A is a cross-

    sectional area.

    10

    -10

    -2 -1

    0 1 2

    5

    -5

    ILn

    AqDnnp0

    V/VT

    Figure 5-3:

    The differential resistance Rd, defined bydIdV

    1

    , is approximately given by VT/I under a

    strong forward bias condition. The diffusion capacitance Cdif of the diode is defined as the voltage

    derivative of the (total) minority carrier charge:

    Cdif ddV

    Q(minority carrier) = Ad

    dV

    q

    0[np(x) np0]dx

    = AqLnnp0

    VTeVVT

    IVT

    n . (5.12)

    7

  • 8/3/2019 Yamamoto Noise

    63/195

    The CR time constant characterized by the differential resistance Rd and the diffusion capacitance

    Cdif is equal to the electron (minority carrier) lifetime n.

    The depletion-layer capacitance Cdep of the diode is defined as the voltage derivative of the

    (total) space charge:

    Cdep ddV

    Q(space charge)

    =2

    WdepA

    =

    q2ND2

    2(VD V) A (5.13)

    Here, Wdep =

    22qND2

    (VD V) is the depletion layer width in the N-layer. The capacitance con-

    tributed by the depletion layer in the p+-layer is neglected. The CR time constant characterized

    by the differential resistance Rd and the depletion layer capacitance Cdep is equal to the thermionic

    emission time te, which we will discuss later in this chapter.

    It will be shown that the thermionic emission time te = CdepRd is a key parameter for deter-

    mining the noise characteristics of a pn junction diode under weak forward bias, while the minority-

    carrier lifetime n = CdifRd is a key parameter for determining the noise characteristics of a pn

    junction diode under strong forward bias. The junction capacitance of a pn junction is determined

    by the depletion-layer capacitance under weak forward bias and by the diffusion capacitance under

    strong forward bias.

    5.1.2 Thermal Diffusion Noise

    When a pn junction is biased by a constant voltage source, the electron (minority carrier)

    densities at x = 0 (edge of the depletion layer) and x = W (p-side metal contact) are constantly

    8

  • 8/3/2019 Yamamoto Noise

    64/195

    fixed to be np0 eV/VT and np0, respectively. The electron density fluctuates, however, between

    x = 0 and x = W due to microscopic electron motion induced by thermal agitation and electron

    generation and recombination processes. In order to keep the boundary conditions at x = 0 and

    x = W and to restore the steady-state electron distribution, the relaxation current pulse flows in

    the entire p+-region between x = 0 and x = W. This relaxation current inside the p+-region results

    in the departure from charge neutrality of this region, which induces the external circuit current.

    If an electron makes a transit over a small distance f between collisions with the lattice, an

    instantaneous current q(t) flows at the two locations x = x

    and x = x

    + f, as shown in Fig. 5-4.

    This instantaneous current creates the departure from the steady-state electron distribution (5.10)

    and triggers the relaxation current in the entire p+-region between x = 0 and x = W, which,

    after a reasonably short time, restores the steady-state electron distribution (5.10). The electron

    distribution deviation n

    (x, t) = n(x, t) np0(x), which results in such a relaxation current in the

    entire region, satisfies the diffusion equation (5.7) and the boundary conditions n

    = 0 at x = 0

    and x = W. The Fourier transform of the diffusion equation (5.7) is given by

    2

    x2N

    (j) =1

    L2N

    (j) , (5.14)

    where N

    (j) is the Fourier transform of n

    (x, t) and

    L2 =L2n

    1 +jn. (5.15)

    The Fourier transform of the relaxation currents i

    1(t) at x = x

    and i

    2(t) at x = x

    + f are

    expressed in terms of the Fourier-transformed electron density deviations N

    1(j) at x = x

    and

    9

  • 8/3/2019 Yamamoto Noise

    65/195

    q(t)

    x = 0 x = Wx' x' + l f

    initial current

    relaxation current

    diffusion current direct return current

    0

    N' ( j)

    N' 1

    N' 2

    Wx

    Figure 5-4:

    N

    2(j) at x = x

    + f:

    I

    1(j) = qDnN

    (j)

    x

    x=x = k1N1(j) , (5.16)I

    2(j) = qDnN

    (j)

    x x=x+f = k2N

    2(j) , (5.17)

    where

    k1 =qDn

    Lcoth

    x

    L

    , (5.18)

    k2 =qDn

    Lcoth

    W x

    L

    . (5.19)

    There are also direct return curre