Messtechnik Chaptersspectros Signal Process
Messtechnik Chaptersspectros Signal Process
G. Jeschke
ETH Zürich, Laboratorium für Physikalische Chemie
[email protected]
HCI F227, Tel. 25702
1 What is spectroscopy? 9
1.1 Spectroscopy as characterization of quantum systems . . . . . 9
1.1.1 A definition of spectroscopy . . . . . . . . . . . . . . . 9
1.1.2 Description of the quantum system . . . . . . . . . . . 10
1.1.3 Description of the sample . . . . . . . . . . . . . . . . 11
1.1.4 Interaction of the sample with the environment . . . . 12
1.1.5 Interaction of the spectrometer with the sample . . . . 14
3 Fourier transform 24
3.1 Integral transforms . . . . . . . . . . . . . . . . . . . . . . . . 24
3.1.1 Fourier series . . . . . . . . . . . . . . . . . . . . . . . 25
3.1.2 Orthonormal systems of functions . . . . . . . . . . . 27
3.2 Definition and properties of the Fourier transform . . . . . . 28
3.2.1 Definition . . . . . . . . . . . . . . . . . . . . . . . . . 28
1
3.2.2 Basic properties . . . . . . . . . . . . . . . . . . . . . 28
3.2.3 Convolution and deconvolution . . . . . . . . . . . . . 31
3.2.4 Correspondences . . . . . . . . . . . . . . . . . . . . . 31
3.3 Representations of complex spectra . . . . . . . . . . . . . . . 34
3.3.1 Absorption and dispersion spectrum . . . . . . . . . . 34
3.3.2 Magnitude spectrum . . . . . . . . . . . . . . . . . . . 35
3.3.3 Power spectrum . . . . . . . . . . . . . . . . . . . . . 36
4 Basics of electronics 37
4.0.4 Energy, power, amplitude, impedance . . . . . . . . . 37
4.0.5 The dB scale . . . . . . . . . . . . . . . . . . . . . . . 39
4.1 Circuit elements and rules . . . . . . . . . . . . . . . . . . . . 40
4.1.1 Circuit elements . . . . . . . . . . . . . . . . . . . . . 40
4.1.2 Rules for combining elements to a circuit . . . . . . . 43
4.2 Filters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.2.1 RC low pass . . . . . . . . . . . . . . . . . . . . . . . . 45
4.2.2 Ideal and real filters . . . . . . . . . . . . . . . . . . . 48
4.2.3 Butterworth filters . . . . . . . . . . . . . . . . . . . . 51
4.2.4 Chebyshev filters . . . . . . . . . . . . . . . . . . . . . 51
5 Noise 53
5.1 Noise sources and types . . . . . . . . . . . . . . . . . . . . . 53
5.1.1 Noise sources and signal-to-noise ratio . . . . . . . . . 53
5.1.2 Thermal noise . . . . . . . . . . . . . . . . . . . . . . 55
5.1.3 Shot noise . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.1.4 1/f noise and Brownian noise . . . . . . . . . . . . . . 57
5.1.5 Phase noise . . . . . . . . . . . . . . . . . . . . . . . . 58
5.2 Mathematics of noise . . . . . . . . . . . . . . . . . . . . . . . 59
5.2.1 Noise spectral density and noise figure . . . . . . . . . 59
5.2.2 Cascaded systems . . . . . . . . . . . . . . . . . . . . 62
5.2.3 Techniques for improving the signal-to-noise ratio . . . 63
2
7 Time-discrete signals 73
7.1 Analog-digital conversion . . . . . . . . . . . . . . . . . . . . 73
7.1.1 Digitization . . . . . . . . . . . . . . . . . . . . . . . . 74
7.1.2 Sampling . . . . . . . . . . . . . . . . . . . . . . . . . 75
7.1.3 Aliasing . . . . . . . . . . . . . . . . . . . . . . . . . . 77
7.1.4 Oversampling . . . . . . . . . . . . . . . . . . . . . . . 78
7.2 Fourier Transformation of time-discrete signals . . . . . . . . 79
7.2.1 Computation of a continuous function in frequency
domain . . . . . . . . . . . . . . . . . . . . . . . . . . 79
7.2.2 Properties of discontinuous FT . . . . . . . . . . . . . 80
7.2.3 Discrete FT . . . . . . . . . . . . . . . . . . . . . . . . 81
7.2.4 Zero filling . . . . . . . . . . . . . . . . . . . . . . . . 83
7.2.5 Fast Fourier transformation . . . . . . . . . . . . . . . 84
7.3 Apodization . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
7.3.1 Effects of limited observation time . . . . . . . . . . . 85
7.3.2 Apodization window . . . . . . . . . . . . . . . . . . . 87
7.3.3 Filtering for optimum signal-to-noise . . . . . . . . . . 88
7.3.4 Filtering for optimum resolution . . . . . . . . . . . . 91
7.4 Network analyzers and spectrum analyzers . . . . . . . . . . . 92
8 Stochastic signals 95
8.1 Characteristics of random variables . . . . . . . . . . . . . . . 95
8.1.1 Probability distribution . . . . . . . . . . . . . . . . . 96
8.1.2 Probability density . . . . . . . . . . . . . . . . . . . . 97
8.1.3 Mean value, variance, standard deviation, moment
analysis . . . . . . . . . . . . . . . . . . . . . . . . . . 97
8.1.4 Gaussian distribution . . . . . . . . . . . . . . . . . . 98
8.1.5 Poisson distribution . . . . . . . . . . . . . . . . . . . 99
8.1.6 Two-dimensional autocorrelation function . . . . . . . 100
8.2 Correlation functions of stationary processes . . . . . . . . . . 101
8.2.1 Stationary processes . . . . . . . . . . . . . . . . . . . 101
8.2.2 Ergodicity theorem . . . . . . . . . . . . . . . . . . . . 101
8.2.3 Autocorrelation function . . . . . . . . . . . . . . . . . 102
8.2.4 Cross correlation functions . . . . . . . . . . . . . . . 103
8.2.5 Correlation functions of periodic signals . . . . . . . . 104
8.2.6 Felgett’s advantage . . . . . . . . . . . . . . . . . . . . 107
3
9 Alternatives to one-dimensional Fourier transformation 108
9.1 Spectral estimators and parameter estimators . . . . . . . . . 108
9.1.1 Asssumptions on the observed system . . . . . . . . . 108
9.1.2 2D spectroscopy . . . . . . . . . . . . . . . . . . . . . 109
9.1.3 Deficiencies of spectrum estimators . . . . . . . . . . . 112
9.1.4 Parameter estimators . . . . . . . . . . . . . . . . . . 113
9.2 Harmonic inversion . . . . . . . . . . . . . . . . . . . . . . . . 114
9.2.1 Linear prediction . . . . . . . . . . . . . . . . . . . . . 115
9.2.2 Singular value decomposition . . . . . . . . . . . . . . 117
9.3 Fredholm integral equations . . . . . . . . . . . . . . . . . . . 119
9.3.1 Fredholm integral equations of the first kind . . . . . . 120
9.3.2 Well-posed and ill-posed problems . . . . . . . . . . . 121
9.4 Regularization . . . . . . . . . . . . . . . . . . . . . . . . . . 122
9.4.1 Tikhonov regularization . . . . . . . . . . . . . . . . . 122
4
10.4.6 Transfer function of FIR systems . . . . . . . . . . . . 142
10.4.7 Transfer function of IIR systems . . . . . . . . . . . . 142
10.4.8 Linear phase systems . . . . . . . . . . . . . . . . . . . 142
10.5 Matlab tools for digital signal processing . . . . . . . . . . . . 145
10.5.1 Useful Matlab functions for system analysis and design 145
10.5.2 Filter design with Matlab . . . . . . . . . . . . . . . . 146
5
Introduction
Scope of this lecture course
This lecture is concerned with spectroscopic measurements and basic pro-
cessing of signals. Only those aspects are treated which are common to
different types of spectroscopy. In fact, many of the topics are also relevant
for a broad range of other measurements in physics and physical chemistry.
The lecture course discusses the procedures that are required to obtain the
most precise, most reliable and least noisy data on a system under study.
Such a general approach implies that the subject matter is rather abstract.
To relate the abstract concepts to spectroscopy, Chapter 1 is devoted to a
general discussion of spectroscopy. This Chapter clarifies what we want to
measure and how the system under study generally responds to excitation.
In Chapter 2 we go to an even higher level of abstraction by treating the
system under study as a black box. This black box receives some inputs
and responds with some outputs. We are interested in only the transfer
function that converts the inputs into outputs, not with the underlying
physics. We ask the question how we can find this function from pairs of
input (excitation) and output (signal). To simplify the treatment we assume
that the system responds linearly.
Both the general behavior of quantum ssytems during spectroscopic experi-
ments and linear response theory call for mathematical methods that relate
time and frequency domain. Such methods are integral transforms, in par-
ticular, the Fourier transform. In Chapter 3 we discuss general properties
of the Fourier transform, which are important for signal analysis.
Signals are detected and partially modified by measurement electronics.
Chapter 4 treats the most important basics of electronics, in particular, how
electronics modifies systems in time and frequency domain. The concept of
signal filtering is introduced. This sets the stage for consideration of noise in
Chapter 5. We consider how noise can be characterized, how it is averaged,
and how it can be reduced or partially supressed. Chapter 6 discusses one
particularly important method for noise reduction, effect modulation with
phase-sensitive detection.
Modern spectroscopy is based on digitization and discretization of signals.
These concepts are treated in Chapter 7, where analog-digital conversion,
sampling issues, and the discrete Fourier transformation are discussed. Fi-
nite observation time of the response may require apodization of the signal
6
by multiplication with a window function, which is introduced as well.
Spectroscopic signals may be related to stochastic processes and are super-
imposed by noise that originates from stochastic processes. To allow for
a quantitative treatment of such situations, Chapter 8 introduces tools for
characterization of stochastic signals, in particular, parameters of random
variables and correlation functions.
The relation between a spectroscopic signal and the underlying physical
information is not necessarily well described by a Fourier transformation. In
some important situations the physical data are related to the spectroscopic
data to a Fredholm integral equation, which may pose a mathematically
ill-posed problem. Such problems can be solved by Tikhonov regularization,
which is introduced in Chapter 9.
Analog filtering and phase-sensitive detection for noise reduction as dis-
cussed in Chapters 4 and 6 are increasingly replaced by equivalent processing
of digital data with digital signal processors. This is discussed in Chapter
10.
Literature
The following textbooks in German language provide good introductions of
the mathematical and theoretical concepts:
[Leon et al.] Fernando Puente Leon, Uwe Kiencke, Holger Jäkel, Signale
und Systeme, Oldenbourg.
Among those, [Meyer] and [Werner] are most easily accessible, whereas [Leon
et al.] is most mathematically strict. The book by [Kiencke & Eger] is geared
at electrical engineers and contains some material that is treated in basic
science courses. On the other hand it also contains some information on
optimization of signal processing electronics, which is missing in the more
abstract textbooks.
7
Many, but not all important concepts are treated in English language in
http://www.analog.com/en/processors-dsp/learning-and-development/
content/scientist engineers guide/fca.html
Acknowledgement
Helpful discussions with and corrections from Dr. Maxim Yulikov, Dr.
Yevhen Polyhach, Andrin Doll, and Rene Tschaggelar are gratefully ac-
knowledged.
8
Chapter 1
What is spectroscopy?
The multi-level quantum system is a small part of the world. The large
rest of the world is the environment of this system. A special part of this
environment is our measurement device- the spectrometer- which we can
manipulate at will and which we can use in turn to manipulate the quan-
tum system. Interaction of the measurement device and of other parts of
the environment with a single instance of the quantum system is solely by
electromagnetic fields, which can be static or oscillatory. In most cases we
work with ensembles of quantum systems and then the temperature, which
we can also manipulate, is an important parameter. Temperature generally
influences the thermodynamic equilibrium state of an ensemble of quantum
systems as well as the interaction of other parts of the environment with
this ensemble.
In this lecture course, the ensemble of quantum systems is termed sample.
The term environment excludes the spectrometer. Our world thus consists
9
heat
Spectrometer
electro-
magnetic fields Environment
at temperature
T
electro-
Sample magnetic
(ensemble of
quantum systems) fields
The eigenvalues i are the energies of the levels. Each level corresponds to
an eigenstate |ii = ψi .
We are interested in the behavior and state of the unperturbed system in a
mean, time-averaged state of the environment. The corresponding Hamil-
tonian is the static Hamiltonian Hb0 with eigenvalues 0,i . We can infer
energy differences between eigenstates by exciting and observing transitions
between eigenstates. The energy of transition (i, j) is
10
where ωij is the angular frequency of the transition, νij the frequency of the
h
transition, h Planck’s quantum of action, and ~ = 2π .
In general, the state of the system is time dependent, Ψ = Ψ(t). Only
eigenstates of H
b are time invariant. The evolution of the system is described
by the time-dependent Schrödinger equation
d i b
Ψ = − HΨ . (1.3)
dt ~
d i hb i
ρb(t) = − H(t), ρb(t) . (1.6)
dt ~
11
and the phase by φij = arctan [< (ρij ) /= (ρij )]. In analogy to Eq. (1.4),
evolution of the coherence conforms to
d 1
Pij (t) = − (Pij (t) − Peq,ij ) , (1.10)
dt T1,ij
12
whereas coherence relaxes according to
d 1
ρij (t) = − ρij (t) . (1.11)
dt T2,ij
We shall see later that T2,ij determines the homogeneous linewidth. Relax-
ation of populations sets a lower limit to the rate constant 1/T2,ij , which is
given by
1 1
≥ . (1.12)
T2,ij 2T1,ij
This effect is termed lifetime broadening of the spectral line. Lifetime broad-
ening is related to Heisenberg’s uncertainty relation for time and energy
∆ (i − j ) ∆t % h.
The rate constant 1/T1,ij contains contributions from coupling of the radia-
tion field to the quantum system and from energy transfer between degrees
of freedom of the environment and the quantum system. The former contri-
bution, which includes spontaneous and stimulated emission dominates in
optical spectroscopy. In magnetic resonance spectroscopy, where frequencies
are much smaller and the strength of coupling to an incoherent radiation
field is exceedingly small, the latter contribution dominates.
An additional contribution to 1/T2,ij arises from fluctuating fields that lead
to stochastic fluctuation of ωij . This in turn leads to stochastic phase fluctu-
ations of subensemble wavefunctions, see Eq. (1.4), and thus to destructive
interference. Consequently, homogeneous linewidths may be larger than
would be expected from lifetime broadening alone.
Interaction of the ensemble with the environment may also lead to transfer
of polarization or coherence between transitions (i, j) and (k, l). Such cross
relaxation is negligible between coherences, unless the two transitions have
the same frequency. If the frequency differs, the coherence is lost due to the
stochastic nature of the transfer.
Within the approximations made in this section, coupling between the sam-
ˆ,
ple and the environment can be described by a relaxation superoperator Γ̂
with element Γij,kl describing relaxation between elements ρij and ρkl of the
density operator. This description by a stochastic Liouville equation,
d i hb i
ˆ {b
ρb(t) = − H b(t) − Γ̂
0, ρ ρ(t) − ρbeq } , (1.13)
d ~
13
Since only ρb is time dependent in Eq. (1.13), this equation can be for-
mally solved. The solution shows that coherence on transition (i, j) evolves
according to
1
ρij (t) = exp − iωij + t ρij (0) . (1.14)
T2,ij
Hence, ωij and T2,ij can be inferred from the time evolution of a signal that
is proportional to coherence on transition (i, j). This is the approach of
time-domain spectroscopy. Alternatively, the spectrum can be measured by
scanning an excitation frequency νexci in the range of expected transition
frequencies and observing absorption or emission. In such frequency-domain
spectroscopy the spectral lines are centered at νij = ωij /(2π) and their width
is related to 1/T2,ij . The information obtained by time- and frequency-
domain spectroscopy is thus equivalent.
14
c = gµB /~ depends on the electron spin g factor and the Bohr magneton
µB , and C
b = Sbx as well as Sb = Sby are spin operators, assuming that the
static field is along z.
The excitation Hamiltonian H
b1 (t) is given in a laboratory-fixed frame, which
relates to the direction of field polarization. The system is not necessarily
quantized along the z axis of this frame. To see, which transitions are driven
by the electromagnetic field, H b1 (t) must be transformed into the eigen-
frame of the static Hamiltonian Hb0 . This transformation can be preceded
by transformation into a rotating or interaction frame, where H
b1 becomes
time-independent. The amplitude of the off-diagonal element Hij of the ma-
trix representation of H
b1 in the eigenframe of H
b0 is the transition amplitude
aij for transition (i, j), D E
aij = i|H
b1 |j . (1.16)
15
resonance, which results in line broadening. Such power broadening distorts
line shapes and may spoil resolution.
Another distortion arises in double resonance experiments when coherence
on transition (i, j) is observed while transition (k, l) is irradiated. If the
irradiation frequency is far off resonance, no transitions (i, j) are driven.
However, a Bloch-Siegert shift,
ωij
∆ωBS = ω12 2 2 , (1.18)
ωij − ωkl
16
Chapter 2
17
2.1.2 Linearity
A simple theoretical treatment is possible if the system is linear. Linearity
is defined by the operator equation
18
2.1.4 Causality
Furthermore, we characterize systems by causality. A system is causal if
and only if its response depends only on the current value and past values
of the input signal, but not on future values of the input signal. Hence, if
x1 (t) = x2 (t) for all times t ≤ t1 , it follows that y1 (t) = y2 (t) for all times
t ≤ t1 . Causality is a general property of systems that we study in science.1
2.1.5 Memory
Systems can be dynamic. The response of a dynamic system depends not
only on the current value of the input signal, but also on past values. In
other words, a dynamic system has a memory. Except for a few cases of
limiting behavior, spectroscopic systems are dynamic.
2.1.6 Stability
Systems can further be stable or instable. A stable systems responds with a
bounded output signal to a bounded input signal. In other words, if there
is a signal amplitude m < ∞ so that |x(t)| < m at all times t, there is also
a signal amplitude M < ∞ so that |y(t)| < M at all times t. An integrator
is an example for an unstable system.
1
At least we assume that and science appears to work well based on this assumption
19
2.2.1 Impulse response
The response g(t) of the system S to a Dirac delta function δ(t) (unit impulse
function, see Box 2.2) is the impulse response. Any input signal can be
expressed as Z ∞
x(t) = x(t − τ )δ(τ )dτ . (2.6)
−∞
Using time invariance, Eq. (2.5), and the superposition principle, Eq. (2.3),
we find
Z ∞
y(t) = S {x(t)} = S x(t − τ )δ(τ )dτ
−∞
Z ∞ Z ∞
= S x(µ)δ(t − µ)dµ = x(τ )S {(δ(t − τ )} dτ
Z ∞ −∞ −∞
where the asterisk operator ∗ denotes the convolution integral. In the first
step we substituted t − τ → µ, which implies the substitution τ → t − µ.2 In
the next step we renamed µ as τ and used the superposition principle. The
results shows that the response to any input signal x(t) is the convolution of
this input signal with the impulse response of the system. In other words,
an LTI system is fully characterized by its impulse response. The meaning
of convolution can be grasped from the graphic representation in Figure 2.2.
* =
Figure 2.2: Convolution of a stick spectrum with a Gaussian line shape function.
20
which follows from the substitution t0 = t − τ . It is also associative,
Hence, if a signal is split, the two parts pass two different systems, and are
then recombined, the impulse response is the sum of the impulse responses
of the two systems.
a c h1
h2
b d
x(t) h1 * h2 y(t) x(t) h1 + h2 y(t)
There is a simple relation between the derivatives of the output signal, the
input signal, and the impulse response
d d d
y(t) = x(t) ∗ g(t) = x(t) ∗ g(t) . (2.11)
dt dt dt
21
Proofs for both propositions can be found in [Leon et al.].
which implies
d
g(t) = a(t) . (2.16)
dt
The impulse response of an LTI system is thus completely determined by the
step response. Hence, the LTI is fully characterized by the step response.
There exists a simple relation between the output signal and the derivatives
of the input signal and of the step response:
d d
y(t) = x(t) ∗ a(t) = x(t) ∗ a(t) . (2.17)
dt dt
Important characteristics of the step response are the rise time trise , the
overshoot, ringing, and the settling time. These characteristics also play a
role in pulse spectroscopies. They are illustrated in Figure 2.4.
22
overshoot
90%
10%
Figure 2.4: Typical step response function and main characteristics. Modified
from Wikimedia Commons.
with a real part R(ω), an imaginary part I(ω), the amplitude spectrum
A(ω), and the phase response function θ(ω). The real part is the absorption
spectrum and the imaginary part the dispersion spectrum.
23
Chapter 3
Fourier transform
where K(t, ν) is the kernel of the transform. This concept can be extended
to the relation of time-domain signals to other variables instead of frequency
ν. Table 3.1 provides an overview of important integral transforms.
24
Note also that integral transformations can be inverted,
Z ∞
y(t) = K −1 (t, ν)F (ν)dν , (3.2)
−∞
25
its overtones sin(2πk/T ) with prefactors
(
1 (−1)(k−1)/2 for n odd
bk = 2 2 . (3.3)
8π k 0 for n even
y(t)
a b c
Figure 3.2: Gibbs phenomenon. The overshoot at a jump discontinuity does not
vanish if a larger number of terms is used for Fourier synthesis of a rectangle wave.
(a) Terms up to k = 10. (b) Terms up to k = 50. (c) Terms up to k = 250.
Adapted from Wikimedia Commons.
1
For the infinite series, the overshoot decays infinitely fast, which ensures that in the
limit the function is perfectly reproduced.
26
In general, the Fourier series can be expressed as
∞ ∞
a0 X 2πkt X 2πkt
y(t) = + ak cos + bk sin . (3.4)
2 T T
k=1 k=1
The real coefficients ak are all zero for odd functions, while the real coeffi-
cients bk are all zero for even functions.
27
3.2 Definition and properties of the Fourier trans-
form
3.2.1 Definition
Only periodic signals can be analyzed with the Fourier series. To analyze
arbitray signals, we may go from a discrete set of basis functions to a contin-
uous distribution of such functions. The sum is thus replaced by an integral
and we have for the Fourier transform (FT)
Z ∞
Y (ω) = y(t)e−iωt dt , (3.9)
−∞
and
y(t) = F −1 {Y (ω)} (3.12)
This is a sufficient, but not necessary condition for the existence of the
Fourier integral.
The FT is linear,
28
complex conjugated spectrum with inverted frequency axis,
F eiω0 t y(t) = Y (ω − ω0 ) .
(3.19)
29
Integration of the time-domain signal can be expressed as
Z t Z t1 Z tn−1
F dt1 dt2 . . . dtn s(tn ) = (iω)−n Y (ω) . (3.22)
−∞ −∞ −∞
The area under y(t) is given by the value Y (0) and the area under Y (ω)
(integral over the whole spectrum) is given by y(0).
Parseval’s theorem
For a pair of time-domain data y1 (t) and Y2 (t) and their FTs Y1 (ω) and
Y2 (ω), the following equation holds:
Z ∞ Z ∞
1
y1 (t)y2∗ (t)dt = Y1 (ω)Y2∗ (ω)dω . (3.23)
−∞ 2π −∞
1
ky(t)k2 = ||Y (ω)| |2 . (3.25)
2π
For the relation between time-domain (t) and frequency domain (ν), the
factor 1/(2π) is dropped,
Riemann-Lebesgue lemma
30
3.2.3 Convolution and deconvolution
Consider the FT of the convolution integral
Z ∞
F {y1 (t) ∗ y2 (t)} = F y1 (τ )y2 (t − τ )dτ
−∞
Z ∞ Z ∞
= e−iωt y1 (τ )y2 (t − τ )dτ dt
−∞ −∞
Z ∞ Z ∞
−iωτ −iω(t−τ )
= y1 (τ )e y2 (t − τ )e dt dτ
−∞ −∞
Z ∞
= y1 (τ )e−iωτ Y2 (ω)dτ = Y1 (ω)Y2 (ω) , (3.28)
−∞
where we have made use of the fact that a double integral does not depend
on the sequence of integration. Hence, the convolution integral in time-
domain transforms to a product in frequency domain Convolution theorem.).
Since the FT can be computed very fast by the Cooley-Tukey algorithm,
convolution can be computed conveniently by FT of the two time-domain
signals, multiplication in frequency domain, and inverse FT.
Furthermore, this property of the FT allows for solving the inverse problem
of deconvolution. Assume that all lines in a spectrum are afflicted by an
additional Gaussian broadening of the same width due to an instrumental
reason. With inverse FT z(t) of the broadening function Z(ω) and the
inverse FT y(t) of the spectrum Y (ω) we can compute a time-domain signal
y(t)
ydeconv (t) = (3.29)
z(t)
3.2.4 Correspondences
Intuitive qualitative analysis of time-domain signals and intuitive qualitative
design of excitation waveforms with desired spectral selectivity are possible
if one knows a few correspondences between functions in time-domain and
frequency domain.
31
Constant offset
1
F {1} = δ(ω) = δ(ν) . (3.30)
2π
∞ ∞
( )
X 1 X 2πk
F δ(t − kT ) = δ(ω − ). (3.31)
T T
k=−∞ k=−∞
y(t) Y(w)
Figure 3.3: Fourier transform of the sampling function. Delta functions in fre-
quency domain are spaced by ω0 = 2π/T .
32
we find
Z ∞ Z T /2
−iωt
Y (ω) = rT (t)e dt = e−iωt dt
−∞ −T /2
1 −iωt T /2 1 −iωT /2
= − e =− e − eiωT /2
iω −T /2 iω
2 ωT sin(ωT /2)
= i sin =T = T sinc (νT ) . (3.33)
iω 2 ωT /2
Exponential decay
Gaussian decay
33
Harmonic functions
1
F {cos(ω0 t)} = [δ(ω + ω0 ) + δ(ω − ω0 )]
2
i
F {sin(ω0 t)} = [δ(ω + ω0 ) − δ(ω − ω0 )] . (3.36)
2
Note that the factor of 2 in the denominator of the last two expressions origi-
nates from integration between −∞ and ∞ and would vanish for integration
of a causal response from 0 to ∞. In that case, however, an imaginary dis-
persion contribution would also arise.
34
and vice versa. The absorption spectrum is the real part and the dispersion
spectrum the imaginary part of the complex spectrum.
1
a 1
b 1
c
Normalized amplitude
0 0 0
Figure 3.4: Effects of dead time in absorption and magnitude spectra. Model
computation for two exponentially damped oscillations with frequencies ν1 = 20
and ν2 = 24 MHz, equal decay time constants of 0.3 µs, and amplitudes a1 = 1 and
a2 = 0.3. (a) Absorption spectrum obtained from time-domain data starting at
time t = 0. (b) Absorption spectrum obtained from time-domain data starting at
time t = 0.8 µs. (c) Magnitude spectrum obtained from time-domain data starting
at time t = 0.8 µs.
35
3.3.3 Power spectrum
The power spectrum is the square of the magnitude spectrum,
36
Chapter 4
Basics of electronics
Eel = V It (4.1)
V2
P = = RI 2 . (4.3)
R
where V and I are the complex voltage and current amplitudes and φ and
ψ are phases, which generally differ for voltage and current. The equivalent
37
to the resistance is the impedance
V
Z= , (4.5)
I
which is a complex quantity, since V and I are complex due to the phase
factors e−iφ and e−iψ . The imaginary part of the impedance is the reactance,
whereas the real part is the resistance.
In the following we assume that the maximal linear dimension dmax and
the minimal duration tmin of the considered electrical processes are in a
range where the speed of propagation of electrical and magnetic fields, the
velocity of light c, can be considered as infinite, i.e. tmin dmax /c. The
corresponding theory of quasi-static currents cannot be applied in the mi-
crowave range of frequencies and generally not to antennas or delay lines.1
It is still applicable in many, but not all cases to radiofrequency circuits.
In the regime of quasi-stationary currents, impedances in series add (Z =
Z1 + Z2 ), while for impedances in parallel we have
−1
1 1
Z= + h parallel circuit i . (4.6)
Z1 Z2
For the series circuit shown in Figure 4.1 Ohm’s law provides the voltage
divider rule Spannungsteiler-
Z2
Vout = Vin . (4.7) formel
Z1 + Z2
Z1
Vin
Z2 Vout
38
The instantaneous power of an alternate current in a two-terminal circuit is
39
by defining a reference point of 1 mW power at 1 dBm, so that
P
(P )dBm = 10 log10 . (4.12)
1 mW
(IEEE)
L
(IEC)
(IEEE)
Figure 4.2: Symbols for selected circuit elements. For resistors and inductors,
both the American (IEEE) and European (IEC) symbol are shown.
40
Resistors
ZR = R . (4.13)
so that a resistance does not introduce a phase shift between voltage and
current.
Inductors
d
V (t) = L I(t) . (4.15)
dt
Since the derivative of a cosine function is the negative sine function (phase
shift -90◦ ), in an ideal inductance circuit the current lags the voltage by
41
ψ − φ = 90◦ . From Eqs. (4.4,4.5,4.15) we can derive
ZL = iωL , (4.16)
Capacitors
The most simple capacitor are two conducting plates separated by air. An
ideal capacitor has infinite resistance for direct currents. Alternate currents
R
can flow, as the charge Q = Idt can be stored in the capacitance C. The
linear relation between voltage and current is
d 1
V (t) = I(t) . (4.17)
dt C
In an ideal capacitance circuit the voltage lags the current by 90◦ , i.e.,
ψ − φ = −90◦ .
In loose terminology it could be said that inductance and capacitance are in
a inverse relationship with each other. More precisely, while an inductance
differentiates current, a capacitance integrates it
Z t
1 Q(t)
V (t) = V (0) + I(t)dt = . (4.18)
0 C C
1
ZC = (4.19)
iωC
42
them by an equivalent circuit diagram 2 .
This approach provides an approximation of the real system by a lumped
element model, which is equivalent to reducing the state space of the real
systems, which is continuous in time and space to a finite state space with
a finite number of parameters. For instance, a loop-gap resonator used
in EPR spectroscopy and in large-volume NMR and magnetic resonance
imaging (MRI) applications can be understood to a good approximation as
an RLC resonance circuit. Not all aspects can be understood in this model,
though. For instance, the spatial distribution of the electric and magnetic
fields in the resonator can only be modeled by renouncing a description by
quasi-static currents.
A R B C R
I1 I -I2 I1 I2 I1
+ I III -
-I3 I
I1 + I2 + I3 = 0
C V1 V2 V1 C V2
- II IV + II
II I1 I2
Figure 4.3: Two-port networks. (A) RC low pass filter with illustration of Kirch-
hoff’s current law for junction I. (B) General scheme for a two-port network. (C)
RC low pass filter.
Kirchhoff’s current law (also Kirchhoff’s junction rule) derives from conser-
vation of charge at any junction in a circuit. The current flowing into any
junction in a circuit must equal the current flowing out of that junction. For
a junction with n branches, it can be formulated as
n
X
Ik = 0 . (4.20)
k=1
2
German: Ersatzschaltbild
43
For instance, for the LR low pass filter circuit shown in Figure 4.3A this law
must hold at junctions I and II, which have three branches each.
Kirchhoff’s voltage law (also Kirchhoff’s mesh rule) derives from conserva-
tion of energy. The directed sum of electrical potential differences around
any closed circuit must be zero. The reason is that a charge that goes full
circle in a closed circuit may not have gained or lost energy. As Kirchhoff’s
voltage law considers only the electrical field, it does not hold true in the
presence of fluctuating magnetic fields. This law thus cannot be applied
without corrections to circuits that contain inductances.
The LR low pass filter shown in Figure 4.3A is a special case of a linear two-
port network (Figure 4.3B). The ports are terminal pairs I/II and III/IV.
They satisfy the condition that the sum of signed currents to the two port
terminals must be zero. A two-port network has an input voltage V1 , an
output voltage V2 , and input current I1 and an output current I2 . These
quantities are related by an impedance matrix
" # " #" #
V1 z11 z12 I1
= (4.21)
V2 z21 z22 I2
If the output port (III/IV) is loaded with a resistance R0 , the first port can
be considered as a linear two-terminal element. Open-circuit operation is
defined as I2 (t) ≡ 0. A two-port network can be characterized by measuring
V2 (t) as a function of V1 (t) or of I1 (t) in such open-circuit operation.
L
I1 I2
I III
V1 R V2
II IV
I1 I2
44
As an example of a linear two-port network we consider the LC low-pass
filter shown in Figure 4.4. The voltage drop of this network is due to the
inductance. With Eq. (4.15) we find
d
V2 (t) = V1 (t) − L I1 (t) . (4.22)
dt
The difference between currents I1 and I2 can be expressed with the current
flow through the resistance R. The voltage across this resistance is V2 . With
Kirchhoff’s current law, Eqs. (4.2) and (4.22) we find
1 L d
I2 (t) = − V1 (t) + 1 + I1 (t) . (4.23)
R R dt
For the loaded LR low pass filter we have the additional equation V2 (t) =
R0 I2 (t). Combining this with Eq. (4.22) we can solve for the relation be-
tween voltage and current at the input port
−1
1 1 d
V1 (t) = + 0 I1 (t) + L I1 (t) . (4.24)
R R dt
The relation between voltage and current at the input port follows directly
from setting the left-hand side of Eq. (4.23) to zero and solving for V1 ,
d
V1 (t) = RI1 (t) + L I1 (t) . (4.26)
dt
4.2 Filters
4.2.1 RC low pass
We now consider the RC low pass filter shown in Figure 4.3C in open-
circuit operation. The current I1 is related to the voltage drop VR over the
resistance,
VR = I1 R . (4.27)
45
voltage law and obtain
V1 = VR + VC , (4.28)
where V2 = VC is the voltage drop over the capacitance. From Eq. (4.17)
we have
dV2 I1
= , (4.29)
dt C
and thus, by substituting VR in Eq. (4.28) with the help of Eqs. (4.27) and
(4.29),
dV2
V1 = V2 + RC . (4.30)
dt
In linear response theory, we can interpret V1 (t) as the input signal x(t) and
V2 (t) as the output signal y(t). Defining the differential operator D b = d,
dt
we thus have
x(t) = 1 + RC D
b y(t) , (4.31)
1
y(t) = {x(t)} = S {x(t)} . (4.32)
1 + RC D
b
The system operator S is thus the inverse of the sum of a unit operator and
the product of the differential operator with resistance and capacitance.3
We now proceed with computing the step response of the RC low pass filter.
By solving Eq. (4.31) for the derivative of y(t), we obtain a differential
equation for the response
dy 1
= (x(t) − y(t)) . (4.33)
dt RC
46
In the stationary state, we must have V2 = V1 = V0 . Hence, the solution
has the form
y(t) = V0 + Ke−t/(RC) . (4.36)
Hence, the RC low pass filter approaches the stationary condition with a
time constant τ = RC. This is the rise time of the RC circuit.4
The formal step response function a(t) corresponds to V0 = 1, hence
d 1 −t/(RC)
g(t) = a(t) = e for t ≥ 0 . (4.39)
dt RC
1 0
0.8
A B
1/Ö2 3 dB
0.6
M(w)
q(w)
-45
0.4
0.2
0 -90
0 1 2 3 4 5 0 1 2 3 4 5
RCw RCw
Figure 4.5: Transfer function (frequency response) of an RC low pass filter. (A)
Amplitude function. (B) Phase response.
47
of the impulse response function,
Z ∞
1
G(ω) = e−[iω+1/(RC)]t dt
RC 0
1 1 1
= =
RC iω + 1/(RC) 1 + RCiω
1 RCω
= −i , (4.40)
1 + R2 C 2 ω 2 1 + R2 C 2 ω 2
where the terms in the last line are the real and imaginary part. Hence, the
magnitude spectrum is
r
1
M (ω) = , (4.41)
1 + R2 C 2 ω 2
1
ωRC = . (4.43)
RC
48
0
pass band 0
-10 A B
Magnitude (dB)
stop band
-20
q(w)
-45
-30
-40
-90
-50
-5 0 5 -5 0 5
log10(RCw) log10(RCw)
Figure 4.6: Bode plot for an RC low pass filter. (A) Magnitude function. (B)
Phase response.
pass and stop behavior, we distinguish low pass (Figure 4.7A), high pass (B),
band pass (C), and band stop (D) filters. Ideal filters cannot be realized
by analog circuits. The reason can be recognized by considering the ideal
band pass filter. Such a filter corresponds to multiplying the signal Y (ω)
in frequency domain by a rectangular function and, thus, to convolution of
the time-domain signal with a sinc function. Since the sinc function extends
to infinity towards both negative and positive times, this would imply that
the signal is delayed for an infinite time in the filter to avoid an unphysical
acausal response.
A B
M(w)
C D
M(w)
w w
Figure 4.7: Idealized magnitude spectra of different filter types. (A) Low pass
filter. (B) High pass filter. (C) Band pass filter. (D) Band stop filter
Filter design is based on a theory for normalized low pass filters. Depending
on the application, different approximations to an ideal filter are aimed for.
49
• Butterworth filters: M (ω) is as flat as possible in the pass band.
• Cauer filters: Defined ripple is accepted both in the stop and pass
band, transition to the stop band is even steeper than with Chebyshev
filters.
• Bessel filters: The group delay should be constant in the pass band, i.e.
different frequency components are delayed by the same time. This
preserves waveform.).
For instance, the RC low pass is a critically damped filter. However, M (ω)
is not particularly flat in the pass band, the transition to the stop band is
not very steep, and the group delay in the pass band differs significantly.
Improvement of the approximation generally requires adding more elements.
This concept can be formalized via the Laplace transform H(s) of the im-
pulse response g(t). For linear, time-invariant (LTI) systems with a finite
number of lumped elements (resistances, inductances, and capacitances), the
function H(s) can be written in rational form (partial fraction expansion)
and written as
This function has m zeroes sz1 . . . szm and n poles sp1 . . . spm , which can
be real numbers or occur as complex conjugate pairs. The number of poles
is the filter order. Filters can be realized not only with passive elements
(RLC), but also with active elements, such as operational amplifiers. In the
latter case, the filter requires an external power source.
50
1 A 0 B
-20
0.8
Magnitude (dB)
-40
0.6 -60
M(w)
-80
0.4
-100
0.2 -120
-140
0
-160
0 1 2 3 4 5 -1 0 1 2 3
w/W log10(w/W)
As seen in Figure 4.8 steepness of the transition from pass band to stop band
and flatness of the frequency response in the pass band increase strongly with
filter order.
51
Figure 4.9: Bode gain plot for a Chebyshev type I filter with order n = 4 and
= 1 (3 dB ripple). Taken from Wikimedia Commons.
52
Chapter 5
Noise
• Signal amplification.
• Environment.
53
receiver and the system itself the excitation noise contributes to the noise
level of the primary output signal.
The level of the wanted signal (ideal response of the quantum system) can be
quantified by the power Psignal and the noise level can be quantified by the
power Pnoise . The quality of the primary output signal is thus characterized
by the signal-to-noise ratio (SNR)
S Psignal A2signal
= = 2 , (5.1)
N Pnoise Anoise
where Asignal and Anoise are the amplitudes of the signal and noise, respec-
tively. The signal-to-noise ratio can be given on the dB scale. In spec-
troscopy, the signal-to-noise ratio is often defined as a ration of amplitudes
of the signal and noise rather than a power ratio. Care must be taken which
definition is adhered to by a given author. The ’spectroscopic’ signal-to-noise
ratio is the square root of the ’electronic’ signal-to-noise ratio.
Usually the primary signal is too weak for direct display or digitization and
is amplified. Such amplification, and indeed any modification of the signal
by electronic components can lead to a decrease in the signal-to-noise ratio
by introducing additional fluctuations.
Finally, the receiver as well as components of the spectrometer that are not
ideally shielded pick up unwanted signals from the environment. Examples
are signals from cell phone networks, radio and TV signals, fluctutations of
the line voltage, and vibrations or fluctuating magnetic and electrical fields
from traffic.
A B
ÖáV 2ñ R
Öái 2ñ
Figure 5.1: Equivalent circuits for the description of thermal noise in resistors.
(A) Noise AC source. (B) Noise current generator.
54
5.1.2 Thermal noise
The noise in any circuit with uniform temperature T can be described by
a noise AC source (Figure 5.1(A)) in series with each resistance R. For a
small frequenccy interval dν we have
hν 1
p(ν) = hν . (5.3)
kB T e kB T − 1
0.8
0.6
p(n)
0.4
0.2
0 5 10 15
log10(n/Hz)
1
i2 = 4kB T p(ν)dν , (5.4)
R
55
or to describe the noise source in terms of available noise power (compare
Eq. (5.1),
1 V2 1 2
Pnoise = = i R
4 R 4
= kB T p(ν)dν . (5.5)
V 2 = 4kB T BR (5.6)
and
Pnoise = kB T B . (5.7)
The center frequency is usually defined as the frequency where M (ν) attains
its maximum. Hence, B corresponds to the width of a rectangular bandpass
filter (brick wall filter) with the same maximum throughput and the same
integrated area of the frequency response function as the actual system.
According to Eq. (5.7), noise power thus does not depend on frequency in
the range where the Planck factor is unity. Such noise is termed white noise.
As an example, consider a resistance of 50 Ω at T = 290 K in a receiver with
a bandwidth of 5 MHz. The mean square noise amplitude is V 2 = 4·10−12
V2 , corresponding to about 2 µV root mean square amplitude of the noise.
The noise power, which does not depend on the resistance, is 2 · 10−14 W or
-107 dBm.
From these considerations we can derive two basic possibilities for improv-
ing SNR. First, we could keep the receiver at low temperature. This is
an expensive way of improving SNR, but is done in radioastronomy and
liquid-state NMR (cryoprobe technology). In both cases the expense of the
additional measurement time needed with an ambient temperature receiver
is higher than the expense for a cryoreceiver. Second, bandwidth of the
56
receiver should be limited to the minimum that is required to detect the
signal of interest. Such bandwidth limitation can usually be achieved at low
expense and is commonly applied in spectroscopy.
where e is the elementary charge. We see that shot noise is also white noise.
57
system the power of 1/f noise equals the power of white noise. Above νc ,
white noise dominates, whereas below νc 1/f noise dominates. For bipolar
transistors, the corner frequency is in the kilohertz range, whereas for the
smallest metal–oxide–semiconductor field-effect transistors (MOSFETs) in
modern integrated circuits, it can reach the Gigahertz range.
-40 -20 0 20 40
n (kHz)
Figure 5.3: Phase noise. Simulation for a normal distribution of phase with stan-
drad deviation of 0.1 rad. A frequency of 10 kHz and a transverse relaxation time
of T2 = 0.2ms were assumed.
58
5.2 Mathematics of noise
5.2.1 Noise spectral density and noise figure
Since white noise frequently dominates and since the power of white noise
is proportional to the effective noise bandwidth B, the noise level is often
characterized by the noise spectral density N0 , with
Pnoise
N0 = hwhite noisei . (5.11)
B
Nth = kB T . (5.12)
Si So
System
Ni No
Figure 5.4: If a signal passes a system, the SNR at the output (So /No ) is generally
different from the SNR at the input (Si /Ni ).
Si /Ni
F = . (5.13)
So /No
This equation can be applied without further thought only if the input noise
is thermal noise at the standard temperature T0 = 290 K, as will become
clear below. If the actual physical temperature is T0 , the noise factor can
59
be expressed by an effective noise temperature of the system,
Te = (F − 1) T0 . (5.14)
The noise factor is often specified on the dB scale (Section 4.0.5) and is then
termed noise figure,
FdB = 10 log10 F . (5.15)
Note that in large parts of the literature, the dimensionless noise factor is
also called noise figure.
Noise figures of uncooled low-power amplifiers used in detection arms of
spectrometers are typically in the range between 1.5 and 2 dB up to fre-
quencies of 6 GHz and between 2 and 5 dB up to 35 GHz. High-power
amplifiers have much higher noise figures, in particular, at high frequencies.
For instance, travelling wave tube amplifiers in EPR spectroscopy have noise
figures of about 15 dB.
Components of a system can be classified as components without gain (at-
tenuators) and components with gain G (amplifiers). Components without
gain are characterized by their attenuation2
Pi
L= . (5.16)
Po
(L − 1)T
F =1+ , (5.17)
T0
ÖáV 2ñ G
60
For amplifiers, the gain is specified as
Po
GdB = 10 log10 . (5.18)
Pi
which leads to
No No
= = FG , (5.22)
Ni Nth
So
As Si = G, we thus recover Eq. (5.13), as long as the input noise is thermal
noise at the standard temperature. Noise factors of amplifiers thus have to
be determined with thermal noise at the standard temperature T0 as the
input by measuring noise power at the output.
61
5.2.2 Cascaded systems
If several systems are cascaded, each of them can be characterized by its
noise factor Fj and gain Gj (Figure 5.6). On order to compute the noise
figure F of the cascade of systems with amplification, we cannot simply
apply Eq. (5.13) to the inputs and outputs of the individual systems. Such
an approach would fail, since after the first amplifier the input noise is no
longer thermal noise.
Hence, we have to compute the noise factor F by considering gains and
added noise in all steps.3 For convenience, we define the noise ratio nj after
the j th stage of the cascade as nj = Nj /Nth , where Nj is the noise power
after this stage and Nth is the thermal noise power. For the first stage, Eq.
(5.22) yields
n1 = F1 G1 . (5.23)
We can now define a noise ratio nadded = Nadded /Nth . Since the two noise
ratios add, we obtain after the second stage
n2 F2 − 1
F = = F1 + (5.26)
G1 G 2 G1
j+1
Y
nj+1 = nj Gj+1 + nadded,j+1 = F (j) Gk + (Fj+1 − 1)Gj+1 (5.27)
k=1
as well as
nj+1
F = F (j+1) = Qj+1 , (5.28)
k=1 Gk
3
This derivation is based on the one in A. van der Ziel, Noise, Prentice Hall, Inc. 1954
62
where F (j) is the noise factor of the cascade after j steps. Combining Eq.
(5.27) and (5.28), we find
Fj+1 − 1
F (j+1) = F (j) + Qj . (5.29)
k=1 Gk
F2 − 1 F3 − 1 Fn − 1
F = F1 + + + ... + . (5.30)
G1 G1 G2 G1 G2 . . . Gn−1
Since amplifier gain factors are large numbers, this sum is usually dominated
by F1 . If several amplifiers are cascaded, the one with the lowest noise figure
should thus be the first one in the chain.
Figure 5.6: Noise figure of amplifiers and amplifier cascades. (A) The noise figure
F = No /(Ni G) of a single amplifier is determined by terminating input with a
resistance R that creates thermal noise Nth and measuring noise power No at the
output. (B) In a cascade, input noise of a given amplifier is output noise of the
previous amplifier rather than thermal noise. Noise added by the first amplifier will
dominate, as it is amplified in all further stages.
From the above considerations it follows that the signal-to-noise ratio im-
proves at low temperatures T of the system and narrow bandwidths B. Since
noise is stochastic, it will also partially average on summation of n individ-
√
ual measurements, Nn = nN , whereas the signal simply adds, Sn = nS.
It follows that
S √ √
∝ n ∝ tm , (5.31)
N
where tm is the time spent for the accumulation of the n measurements.
With respect to 1/f noise and slow signal drifts, it is advantageous to mea-
sure at frequencies larger than 1 kHz, or, more generally, larger than the
63
corner frequency, if the quantum system permits. Whether or not the quan-
tum system permits this, may depend on relaxation times and resonance
frequencies.
To reduce spectrometer noise, components can be cooled. According to Friis’
formula, cooling the first detection amplifier is usually sufficient. Effect
modulation with phase sensitive detection can reduce bandwidth and ist
treated in the next Chapter.
Environmental noise can be reduced by shielding of components against ex-
ternal electromagnetic fields. In some cases, the external fields can also
be measured and compensated for. In NMR spectroscopy, slow drifts of the
magnetic field are compensated for by a spin lock, while in EPR spectroscopy
a field-frequency lock can offset field drifts and drifts in the resonator fre-
quency. Effect modulation also reduces the influence of environmental noise
by limiting detection bandwidth.
Impedance matching
64
microwave resonators, coupling between transmission line and resonator is
matched by varzing the size of a coupling hole (iris coupling) or the position
of an antenna with respect to the resonator.
65
Chapter 6
carrier
66
Coherent demodulation is used if a carrier with the same frequency and
phase as used in modulation is available at the receiver. If that is not the
case, incoherent demodulation can be used, which is realized in the simplest
case by a diode and a low-pass filter.
Frequency modulation
A B C
Figure 6.2: Modulation types. (A) Carrier (black) and signal (blue). (B)
Frequency-modulated signal (∆ν = 0.05νc ). (C) Amplitude-modulated signal
(black line) with m = 0.8 with envelopes (dotted blue lines) corresponding to the
signal shape.
67
Amplitude modulation
1
c(t)·x(t) = cos(2πνc t) cos(2πνs t) = [cos(2π(νc + νs )t) + cos(2π(νc − νs )t)] .
2
(6.2)
In other words, sidebands are generated at the carrier frequency plus/minus
the signal frequency.
In practice, these sidebands are admixed to a pure carrier oscillation
m m
y(t) = Ac cos(2πνc t) + Ac cos [2π(νc + νs )t] + Ac cos [2π(νc − νs )t] ,
2 2
(6.3)
where m is the modulation index. Distortionless incoherent demodulation is
possible only for 0 < m ≤ 1.
From Figure 6.3 it is clear that one sideband carries the whole information.
With more involved modulation circuits and receivers, double sideband sup-
pressed carrier (DSBSC) and single sideband (SSB) modulation are possible.
carrier
bandwidth
0 ns nc-ns nc nc+ns n
68
A B D
Vi VC VD Vo C E
69
Io
Vi
Figure 6.5: Signal distortion by a non-linear characteristic curve. The grey input
signal is a pure cosine wave. Due to saturation in the characetristic curve, the blue
output signal is damped near its maximum amplitude.
to the PSD. The modulated laser irradiation will cause sample fluorescence
that is modulated by the same frequency, whereas the phase may differ.
This fluorescence signal is detected by a photo-detector, such as a photo-
diode or CCD camera. The signal from the sample at the output of the
70
photodetector has the form
The mixer multiplies the signal input with the reference signal
Vs,0 Vr,0
Vm (t) = cos (ωr t + φs ) cos (ωr t + φ)
V0
1 Vs,0 Vr,0
= [cos (2ωr t + φs + φ) + cos (φ − φs )] . (6.6)
2 V0
The low-pass filter eliminates the high-frequency term and any noise at
frequencies above the bandwidth of the signal from the sample. This signal
may be slowly oscillating, i.e. φs may be time-dependent at frequencies
much lower than ωr . We find for the output voltage,
1 Vs,0 Vr,0
Vo (t) = cos (φ − φs ) . (6.7)
2 V0
Since V0 is much smaller than Vr,0 the PSD acts as an amplifier. The output
π
signal is maximal for φ = φs and vanishes for φs = φ ± 2. To detect
the maximum signal, the reference signal can be properly phase shifted.
Alternatively, the reference signal can be supplied in quadrature (original
reference signal and reference signal phase shifted by π/2) and the magnitude
of the two quadrature components can be measured.
lock-in amplifier
signal signal
low-pass filter
input output
reference
71
6.2.1 Digital PSD
Modern PSDs are based on digital signal processors. The input signal and
reference signal are first digitized and discretized by a sufficiently fast analog-
digital converter (ADC) (Chapter 7). Phase-shifting, multiplication, and
filtering can then all be performed by a digital signal processor (Section
10.1.1), which adds flexibility and allows for realizing very high-order filters.
If necessary, the output signal is converted to an analog signal by a digital
analog converter (DAC). However, in many cases the digital signal is further
processed.
72
Chapter 7
Time-discrete signals
73
(S/s). An n-bit ADC discretizes the signal with a resolution of 2n possible
values over the input range. For instance, a 3-bit ADC encodes the signal
level as one of eight numbers, as visualized in Figure 7.1. At the high end,
compromises must be made between resolution and sampling rate. As of
October 2011, the fastest ADC is a prototype for development applications
with a sampling rate of 56 Gs/s at 8 bit resolution (Fujitsu). At a resolution
of 12 bit, which is sufficient for most applications, devices with sampling
rates up to 3.6 GS/s (National Semiconductor) are commercially available.
7.1.1 Digitization
The vertical resolution determines the round-off error in discretization. Such
discretization corresponds to a quantization of the signal amplitude. As is
apparent from Figure 7.1 and Figure 7.2B, information on the exact shape
of the signal is lost. If the signal is constant, this loss in information cannot
be recovered by averaging. The deviation of the discretized signal from the
true signal is called quantization noise. Quantisierungs-
A different situation arises if the signal is superimposed by white noise with rauschen
an amplitude that exceeds the resolution, i.e. the change in signal level
corresponding to the least significant bit (LSB). Such a signal is shown in
Figure 7.2C. If such a signal is repeatedly sampled and the digital data are
averaged, quantization noise averages along with the white noise (Figure
7.2C). This is because the noise can push the signal above the next step or
below the previous step and the probability for such events depends on the
exact value of the signal.
Accordingly, the vertical resolution of the ADC should be selected such that
in typical measurements the noise exceeds 1 LSB. If the signal in a single
scan has a very large signal-to-noise ratio S/N> 2n , quality of the data is
limited by the ADC resolution. In such cases it may be advantageous to
switch off the first low-noise amplifier in the receiver. This will drastically
reduce S/N, which can be tolerated in such situations, if signal acquisition
can be repeated at a reasonable rate. With the low-noise amplifier switched
off and noise exceeding 1 LSB, signal quality can be improved to the required
level by averaging a sufficient number of scans.
74
1 1
A B
Normalized signal amplitude
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
1 1
C D
Normalized signal amplitude
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
time (a.u.) time (a.u.)
7.1.2 Sampling
The discretized or sampled signal s is a vector of signal values given at
the sampling points. With the sampling rate νs and the sampling period
T = ∆t = 1/νs , sampling is defined by
sk = y(kT ) k = 1 . . . N , (7.1)
75
sampling period fulfills the sampling theorem,
π 1
T < = . (7.2)
ωmax 2νmax
This equation can be derived by expressing s(t) in a Fourier series, but the
derivation is lengthy.
1 1
A B
Normalized signal amplitude
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
time (a.u.) time (a.u.)
Figure 7.3: Sampling and signal reconstruction (simulation). The signal has two
frequency components ν1 = 5 and ν2 = 2.9 and is sampled with a Nyquist frequency
of 8. (A) Original signal (black), sampling times (blue lines), and sampled signal
points sk (blue circles). (B) Original signal (black dashed line), sampled signal
points Sk (blue circles) and signal reconstructed by Eq. (7.3) from the limited
number of sampled points (red line).
Sampling and signal reconstruction are illustrated in Figure 7.3. With fre-
quency components of 2.9 and 5, a Nyquist frequency of 8 is amply sufficient.
Nevertheless, signal reconstruction is not perfect in this case as the sampled
signal is known not from t = −∞ to t = ∞, but only in the range from 0 to
76
1. This leads to deviations between the reconstructed (red line) and original
signal (dashed line).
7.1.3 Aliasing
We now consider the consequence of sampling in frequency domain. The
discretized signal sd is given by
∞
X
sd = y(t)δ(t − nT ) . (7.4)
n=−∞
1
T < (7.6)
2νmax
must be fulfilled. Only in that case all copies are separated and the Fourier
1 1
transform of the discretized signal in the range (−νNyq , νNyq ) = − 2T , 2T is
exactly equal to S(ω).
If the time step T = ∆t is taken to long, the copies overlap. Frequencies
with absolute values larger than the Nyquist frequency are folded back into
the accessible frequency range and lead to spurious peaks. This phenomenon
is called aliasing and should be avoided. To avoid unpleasant surprises, the
signal can be bandwidth limited by an appropriate low pass filter before
discretization.
77
0.4
4 a
0.2 3
2
0
1
-0.2
0
0 1 2 3 4 5 0 2 4 6 8 10 12
0.4
4 b
0.2 3
2
0
1
-0.2
0
0 1 2 3 4 5 0 2 4 6 8 10 12
0.4
4 c
0.2 3
0 2
1
-0.2
0
-0.4
0 1 2 3 4 5 0 2 4 6 8 10 12
t (ms) n (MHz)
7.1.4 Oversampling
If the ADC can sample faster than is required by the bandwidth limitation of
the signal, signal quality can be improved by using a time step ∆t = T that
is shorter than required by νmax . First, such oversampling guards against
aliasing of noise with frequencies larger than νmax . Such noise may exist
even if the signal has passed an analog low pass filter, since suppression
of signal beyond the cutoff frequency by the filter is not perfect. Second,
78
oversampling reduces the influence of quantization noise in the spectrum,
since digitization noise also has frequency components beyond νmax . This
effect is illustrated in Fig. 7.4. It is the reason for the use of oversampling
techniques in music players.
sk = y(kT ) . (7.7)
We first consider the case where the signal is discrete, but known throughout
all history, presence, and future. This signal can be expressed as a time-
continuous function y∗ (t) by multiplication of the time-continuous signal
with a series of Dirac delta functions,
∞
X
y∗ (t) = y(t) · δ(t − kT ) . (7.8)
k=−∞
Using Eq. (3.9), we thus find for the FT (spectrum) of the discontinuous
(time discrete) signal
Z ∞ Z ∞ ∞
X
−iωt
Y∗ (ω) = y∗ (t)e dt = y(t)δ(t − kT )e−iωt dt
−∞ −∞ k=−∞
∞
X Z ∞ ∞
X
−iωt
= y(t)δ(t − kT )e dt = y(kT )e−iωkT
k=−∞ −∞ k=−∞
X∞
= sk e−iωkT . (7.9)
k=−∞
79
discrete signal is given by
Z π
1
sk = Y∗ (Ω)eiΩkd Ω , (7.10)
2π −π
Linearity
Shift theorem
Convolution theorem
n o
(A) (B)
Fd s(A)
n ∗ s (B)
n = Sk Sk , (7.13)
Power theorem
∞ Z π/T
X T 2
|sn |2 = S(eiωT ) dω . (7.15)
n=−∞
2π −π/T
80
7.2.3 Discrete FT
In most cases one is interested in a spectrum that is also discrete, as the
spectrum is further processed and analyzed by a digital computer. To find
out how the spectrum should be discretized, we note that in any practical
situation only a finite number of the sk in Eq. (7.9) is available. In par-
ticular, we do not know the signal at negative times k < 0 and we don’t
know the signal after the end of our observation time tmax = N T . For the
observation frequency or frequency increment, we define
1 1 νs
∆ν = = = . (7.16)
tmax NT N
In other words, the spectrum has the same number of points as the time-
domain signal.
This discrete spectrum can be obtained by the Discrete Fourier transfor-
mation (DFT), which results from defining a normalized discrete frequency
domain with νn = n∆ν and
The sum over the Ωn substitutes for the integral over ω in Eq. (7.9). This
normalization of angular frequency domain has the advantage that the com-
putation algorithm becomes independent of the actual time axis. Time
vanishes from the equation and is replaced by the running index,
N
X −1
Sn = sk e−i2πkn/N n = 0, 1, . . . N − 1 . (7.18)
k=0
Remarks:
81
discrete spectrum, so that the integral of the spectrum is twice as large
as it should be. To avoid this, s0 needs to be divided by two before
DFT.
• If the input data sk are real, the spectrum is an even function. In that
case no shift is required. The spectrum and frequency axis can simply
be cut after l points for N = 2l and after l + 1 points for N = 2l + 1.
The Matlab function dft, programmed for this lecture course and available
on the lecture course homepage, computes the spectrum and frequency axis
from a given time-domain signal and time axis, taking these considerations
into account.
DFT can be written as a matrix equation
~ = W~s ,
S (7.19)
−1
Wkn = ei2πkn/N . (7.22)
82
With W = e−i2π/N , the matrix assumes the form
1 1 1 1 ...
1 W W2 W3 ...
W=
1 W2 W4 W6 .
... (7.23)
... ... ... ... ...
83
amplitudes. For certain display and analysis purposes extended zero filling
can make sense. However, on should always be aware that the true resolution
1
si not better than ∆ν = 2tmax .
By default, the Matlab function dft applies zero filling to 2N . This behavior
can be changed by specifying in options.zerofilling either no zero filling
(options.zerofilling=0) or extended zero-filling (options.zerofilling
= M, where M specifies the number of points after zero filling and must be
larger than N ).
Reduction in number of multiplications
1200
1000
800
600
400
200
0
0 5000 10000 15000
N
Figure 7.5: Reduction factor of the number of multiplications between FFT and
DFT. Already for n = 16384 points, a factor of more than 1000 is saved in compu-
tation time.
84
This advantage is the main reason why much of digital data processing is
based on data sets whose dimensions are powers of 2. Despite the increase in
computer efficiency, the advantage of the FFT algorithm is still important,
in particular in real-time data processing and for multi-dimensional FT.
7.3 Apodization
7.3.1 Effects of limited observation time
Limited resolution
The limited acquisition time leads to limited resolution. This effect can be
understood semi-quantitatively from Eq. (3.35). If a signal exhibits Gaus-
sian decay with time-domain variance σt2 = 1/(2a), the spectrum will exhibit
an angular frequency-domain variance σω2 = 2a and thus a frequency-domain
variance σν2 = a/(2π 2 ). Hence, the product of the standard deviations is
σt σν = 1/(2π) and the uncertainty of the frequency is σν = 1/(2πσt ).
If the signal is not decayed at maximum observation time tmax , no theoretical
resolution limit is set by the observation time. For instance, a signal that is a
linear combination of a small number n of exponentially damped harmonics
is fully characterized by 2n parameters, namely the n frequencies νij and
n decay time constants T2,ij . If a noise-free signal of that type is sampled
according to the Nyquist criterion with at least N = 2n points, the full
information is available. In practice, a few more points need to be acquired
to reduce influence of noise. In that case frequencies and relaxation times
can be determined without an uncertainty that is related to tmax , if data
analysis techniques other than FT are applied (see Section 9.2). However,
such approaches break down if no theoretical model for the time-domain
signal can be assumed or if n becomes too large.
Truncation artifacts
85
1 1 ms
a 50
b
0.5 40
30
0
20
-0.5 10
0
-1
0 2 4 6 8 -20 -10 0 10 20
t (ms) n (MHz)
20
30
c d
25
15
20
15 10
10
5
5
0 0
Figure 7.6: Truncation artifacts and apodization. (a) Real part of a signal y(t) =
e(i2πν−1/T2 )t with ν = 5 MHz and T2 = 1 µs. The vertical dashed line indicates
truncation at tmax = 1 µs. The red curve is a Hamming apodization function.
(b) Spectrum obtained by DFT after zero-filling from the complete signal with
tmax = 8 µs. (c) Spectrum obtained by DFT of the signal truncated at tmax = 1 µs
after zero-filling (black) and without zero-filling (green). (d) Spectrum obtained
from the signal truncated at tmax = 1 µs by multiplication with the Hamming
window (black dashed) or a Dolph-Chebyshev window with 50 dB ripple suppresion
(red), zero-filling, and subsequent DFT.
which corresponds to
(
y(t) for t ≤ tmax
ytrunc (t) = . (7.25)
0 for t > tmax
sin (ωtmax )
Ytrunc (ω) = Y (ω) ∗ 2tmax . (7.26)
ωtmax
86
Figure 7.6C. Although the ripples are not apparent if DFT is applied with-
out zero filling, the spectral line shape is still distorted. In Figure 7.6C
this distortion is apparent as an asymmetry and some negative intensity in
the high-frequency wing of the line in the green spectrum. Such trunca-
tion artifacts for large signals can conceal small signals and should thus be
avoided.
with
1 α
β = cosh arccosh (10 ) , (7.28)
M
where the ripple level in dB is given by −20α. The required window func-
tion in time domain w(t) can be obained by inverse DFT. In Matlab, the
Signal Processing Toolbox provides a function chebwin that computes the
Dolph-Chebyshev window in time domain for given number of points and
ripple attenuation. The Matlab function dft.m for DFT with pre-processing,
which was written for this lecture course, allows for using Dolph-Chebyshev
apodization if the Signal Processing Toolbox is available. Ripple suppression
in dB (as a positive number) can be supplied in variable options.ripple.
As is seen in Figure 7.6D, the lineshape with Dolph-Chebyshev apodization
(red) with 50 dB ripple suppression (the default in dft.m) is symmetric and
has low ripples. Line broadening is moderate, as is apparent from compari-
87
son with Figure 7.6B.
Hamming window
As is seen by comparing the dashed black line (Hamming window) with the
red line (Dolph-Chebyshev window) in Figure 7.6D, the Hamming window
is a good, simple substitute for a Dolph-Chebyshev window in many situa-
tions. In dft.m the Hamming window is implemented independently from
the Signal Processing Toolbox.
Kaiser window
The Hamming window does not allow for adjusting the compromise between
ripple suppression and line broadening. Such adjustment is possible with the
Kaiser window h p i
I0 πα 1 − t2 /t2max
w(t) = , (7.30)
I0 (πα)
where I0 is the zero-order modified Bessel function of the first kind and α de-
termines the trade-off between ripple suppression and broadening. In dft.m
the Kaiser window is implemented independently from the Signal Processing
Toolbox with a default α = 2 which provides a similar compromise between
ripple and line broadening as the Hamming window. The parameter α can
be supplied in variable options.alpha.
88
1
a 50
b
0.5 40
30
0
20
-0.5
10
0
-1
0 2 4 6 8 -20 -10 0 10 20
t (ms) n (MHz)
50
c 25
1 d
40 20 0.5
0
30 15 -0.5
-1
20 10 0 2 4 6 8
t (ms)
10 5
0 0
clear for a case where the wanted signal has decayed below the
noise level
at
t
time tdecay < tmax . Multiplication of y(t) by a box function Π 2tdecay will
set the time-domain signal to zero at times t > tdecay and will not affect the
signal at t ≤ tdecay . This operation will suppress more noise than signal and
will thus increase signal-to-noise ratio. In frequency domain, such deliberate
truncation of the signal corresponds to smoothing by convolution with a sinc
function. The effect of such multiplication with a box function is illustrated
in Figure 7.7C.
89
Matched filter
We now consider that the signal decays with an envelope function ye (t). For
damped harmonic oscillations yk (t) = exp(iωk t) exp(−t/T2,k ) this envelope
function is ye (t) = exp(−t/T2,k ). Since noise level is time independent, the
lokal signal-to-noise ratio at time t is proportional to ye (t). By multiplying
the signal once more with ye (t), each point of the time domain signal is
weighted with its own sensitivity (local signal-to-noise ratio), so that points
with larger signal-to-noise ratio dominate in the time-domain signal. The
window function
wmatched (t) = ye (t) (7.31)
90
7.3.4 Filtering for optimum resolution
Fourier self deconvolution
a b
Lorentz-Gauss transformation
91
wings can lead to peak-like features (Figure 7.8a) that are hard to distinguish
from true, weaker peaks. The following window function achieves a Lorentz-
Gauss transformation
2 t2 /2
wLG (t) = et/τ −σ . (7.34)
Sinebell apodization
For φ = 0 the first point of the time-domain signal is set to zero, which im-
plies zero integral of the spectrum and thus negative ”line feet”. This effect
is reduced by introducing the phase shift φ. Larger phase shifts correspond
to less resolution enhancement, less reduction of the signal-to-noise ratio and
less negative intensity. Examples for the effect of sinebell transformations
with zero and 10◦ phase shift are given in Figure 7.9b,c.
92
a d
b e
c f
-20 -10 0 10 20 -20 -10 0 10 20
n (Hz) n (Hz)
the system, the transmitted wave can be attenuated or amplified and phase
shifted. Most modern network analyzers are vector analyzers, which mea-
sure the complex signal and can thus characterize not only amplitude, but
also the phase shift and in many cases the group delay. Network analyzers
provide the test signal and measure the response.
Spectrum analyzers characterize time-domain signals in terms of their spec-
trum. Analog spectrum analyzers use the super heterodyne principle. A ref-
93
erence frequency is scanned linearly by supplying a sawtooth control voltage
to a local oscillator. The signal to be characterized is mixed with the ref-
erence signal of variable frequency and then filtered by a narrow bandpass.
The sawtooth control voltage provides the horizontal axis of an oscilloscope
picture, while the amplitude of the filtered signal provides the vertical axis.
Nowadays digital spectrum analyzers are more common. In these devices,
the signal is subjected to DFT, using the FFT algorithm. Spectrum ana-
lyzers for bandwidth-limited high frequency signals are hybrid FFT-super
heterodyne analyzers, which use downconversion with a reference signal to
allow for longer sampling intervals. This makes FFT more efficient, as it
limits the number of points. Real-time spectrum analyzers process different
overlapping parts of the incoming signal in parallel and superimpose the re-
sulting spectra. Modern digital oscilloscopes usually have a simple spectrum
analyzer function without downconversion.
Test signal
System Transmission
Reflection
94
Chapter 8
Stochastic signals
80
60
t0 a b
2
12
40
amplitude (a.u.)
log (M)
1.5 8
20
0 4
1
-20 0
Figure 8.1: Stochastic signal (Brownian noise). (a) The time trace was simulated
by integrating white noise, which was in turn simulated by normally distributed
random numbers (commands randn and cumsum in Matlab). (b) Magnitude spec-
trum of the stochastic signal obtained by DFT. The inset is a double logarithmic
plot.
95
stochastic signals. As an example, consider Brownian noise, which is the
integral of white noise (Figure 8.1a). As white noise has a constant mean
square amplitude as a function of frequency, Brownian noise has a 1/ω 2
dependence of the mean square amplitude (power spectrum). This can be
recognized by FT (Figure 8.1b) and can be seen particularly well in a dou-
ble logarithmic plot (inset). In the case at hand, the slope in the double
logarithmic plot of the magnitude spectrum is -1, as the amplitude has a
1/ω dependence if the power has a 1/ω 2 dependence. Deviations at very
low frequencies are due to the limited observation time.
If we measure the same random variable several times at time t0 after ini-
tializing (triggering) our experiment, we will obtain different values (real-
izations) x1 (t0 ), x2 (t0 ), x3 (t0 ) . . . of the random variable. If we measure (or
simulate) the Brownian noise trace several times, it will look different each
time. However, the 1/ω dependence of the amplitude will be a stable charac-
teristic. In the following, we summarize the most important characteristics
of random variables as they are used in probabilistic theory.
F (−∞) = 0
F (∞) = 1 . (8.2)
96
1
F(x)
a b
p(x)
6
0 1 2 3 4 5 6 0 1 2 3 4 5 6
x x
Figure 8.2: Probability distribution function F (x) (a) and probability density
function p(x) (b) for an ideal dice.
dF (x)
p(x) = . (8.3)
dx
The probability density function for an ideal dice is plotted in Figure 8.2b.
The intergal of the probability density function is unity,
Z ∞
p(x)dx = 1 . (8.4)
−∞
97
The zeroth moment of a function f (x) is thus
Z ∞
M0 = f (x)dx . (8.7)
−∞
In moment analysis, hxi is the first moment M1 ). For noise, we have hxi = 0.
The width of the distribution is characterized by the variance σ 2 , which is
the square of the standard deviation σ. This is computed as
Z ∞
2
σ = (x − hxi) p(x)dx . (8.9)
−∞
σ 2 = M2 − M12 . (8.10)
The variance is also the second central moment, with central moments µk
being defined as
Z ∞
µk = (x − hxi)k f (x)dx k = 2, 3, . . . . (8.11)
−∞
∞ 3
x − hxi
Z
µ3
γ1 = p(x)dx = . (8.12)
−∞ σ σ3
98
The condition of identical distributions of the random variables can be sub-
stituted by several other, weaker assumptions. This situation is often en-
countered in physical sciences. White noise, for instance thermal noise at
low frequencies, conforms to a normal distribution, as does the error of a
measurement that is performed repeatedly under identical conditions.
Note however that in many cases where normal distributions are assumed,
the conditions of the central limit theorem are not fulfilled or at least are
not known to be fulfilled. Whenever possible, the hypothesis of a normal
distribution should be tested before it is used in data analysis. For instance,
the skewness of the normal distribution is zero. Analysis in terms of a normal
distribution will thus conceal any asymmetry of the probability distribution
with respect to the mean value.
The Gaussian function cannot be integrated in closed form. If the proba-
bility density function is a Gaussian function, the probability distribution
function is related to the error function
(x−hxi)/σ
x − hxi
Z
1 2 1
F (x) = √ e−t dt = 1 + erf √ , (8.14)
2π −∞ 2 2σ
with Z y
2 2
erf(y) = √ e−t dt . (8.15)
π 0
e−rt (rt)k
p(k, r, t) = . (8.16)
k!
e−λ λk
p(k, λ) = (8.17)
k!
99
with the property
∞
X
p(k) = 1 . (8.18)
k=0
1
a 0.12 b
0.8
0.08
Sp(k)
0.6
p(k)
0.4
0.04
0.2
0 0
0 5 10 15 20 25 0 5 10 15 20 25
k k
Figure 8.3: Poisson distribution with λ = 10. (a) Sum of all discrete probabilities
from p(0) to p(k), corresponding to the probability distribution F (x) of a continuous
variable. (b) Discrete probability density function p(k) (black dots) and Gaussian
function corresponding to Eq. (8.19) (blue line).
This function tells us whether the signal at time t2 is correlated to the one
at time t1 (positive value), anti-correlated (negative value), or uncorrelated
(zero). The amplitude of the function quantifies the strength of the correla-
tion.
100
8.2 Correlation functions of stationary processes
8.2.1 Stationary processes
We now assume that we repeatedly measure the random variable x at a time
t0 after initializing the experiment. We shall find a probability distribution
F (x) and probability density p(x) at this given time t0 . If these functions
are independent of our choice of t0 , i.e., invariant under time translation,
the process is stationary. For instance, the process that creates thermal
noise is stationary if temperature was constant for a sufficiently long time.
Processes may be stationary at some times during our experiment, but not
at other times.
Stationary processes have time-independent mean value and variance. The
autocorrelation function depends only on the difference τ = |t1 − t2 | of the
two observation times. Therefore, the autocorrelation function is a one-
dimensional function Rxx (τ ) and it suffices to analyze it for τ ≥ 0.
101
8.2.3 Autocorrelation function
The autocorrelation function
Z T
1
Rxx (τ ) = hx(t)x(t + τ )i = lim x(t)x(t + τ )dt = x(t)x(t + τ )
T →∞ 2T −T
(8.22)
of an ergodic process is an even function,
∗ (τ ).
If x(t) is complex, Rxx (τ ) is hermitian, Rxx (−τ ) = Rxx
At t = 0 the autocorrelation functions assumes the mean square value of
x(t),
Z T
2 1
Rxx (0) = hx i = lim x2 (t)dt . (8.24)
T →∞ 2T −T
In particular, for signal voltages x(t) = V (t), the value of the autocorrelation
function at t = 0 is proportional to the average power of the signal,
Z T
1 1
P̄ = lim V 2 (t)dt . (8.25)
R T →∞ T 0
In other words, a box function with amplitude Rxx (0) extending from τ = 0
to τ = τc has the same area as the autocorrelation function. If the auto-
correlation function is an exponential decay, τc is the time constant of this
decay.
The power spectral density (power spectrum) of a signal is the FT of its
autocorrelation function (Wiener-Khinchin theorem)
102
spectra density,
Rxx (t) = F −1 {Sxx (ω)} . (8.29)
Since white noise has uniform spectral density, Eq. (8.29) predicts that the
autocorrelation function of white noise has a peak at τ = 0 and is identical
zero everywhere else. For a finite-time signal, the autocorrelation function
looks noisy for τ > 0.
For a time-discrete signal y(tk ), the autocorrelation function can be effi-
ciently computed by DFT to obtain Y (ωl ), computing the product Z(ωl ) =
Y (ωl )Y ∗ (ωl ) and subjecting this product to inverse DFT. This algorithm
is based on the convolution theorem and the fact that the autocorrelation
function is the convolution of the signal with itself.
Note that this is again a convolution integral, which can be computed ef-
ficiently for time-discrete signals by DFT of both functions to frequency
domain, multiplication of one spectrum with the complex conjugate of the
other spectrum, and inverse DFT. The cross correlation function fulfils
∗
Rxy (τ ) = Ryx (−τ ) . (8.31)
Cross correlation functions are useful for instance for characterizing the cor-
relation between input and output of a device or circuit. As a particu-
larly simple example, we consider the voltage divider shown in Figure 4.1
with x(t) = Vin and y(t) = Vout with purely real impedance Z1 = R1
and Z2 = R2 . We assume an autocorrelation function of the input signal
Rxx (τ ) = σ 2 e−k|τ | . Since
R2
y(t) = x(t) , (8.32)
R1 + R2
we have
R22
y(t)y(t + τ ) = x(t)x(t + τ ) , (8.33)
(R1 + R2 )2
103
so that the autocorrelation function of y(t) is
R22 R22
Ryy (τ ) = Rxx (τ ) = σ 2 e−k|τ | . (8.34)
(R1 + R2 )2 (R1 + R2 )2
R2 R2
Rxy (τ ) = Rxx (τ ) = σ 2 e−k|τ | . (8.35)
R1 + R2 R1 + R2
In this computation we have neglected that the resistors add noise to the
signal. Thus, the result is only valid for the case where the input signal is
much larger than the added thermal noise.
1 1
x(t)x(t + τ ) = A2 cos(ω0 τ ) + A2 cos(2ω0 t + ω0 τ ) . (8.37)
2 2
A2
Rxx (τ ) = cos(ω0 τ ) . (8.38)
2
For
y(t) = B sin(ω0 t) , (8.39)
we have
1 1
y(t)y(t + τ ) = B 2 cos(ω0 τ ) − B 2 sin(2ω0 t + ω0 τ ) . (8.40)
2 2
and find
B2
Ryy (τ ) = cos(ω0 τ ) . (8.41)
2
We find average powers Rxx (0) = A2 /2 and Ryy (0) = B 2 /2 and in both
cases the period of the autocorrelation function is the same as the one of
104
the input function x(t).
This is a general property, as can be shown by expressing an arbitrary
function x(t) with period T by a Fourier series. With Eq. (3.4) we find
∞ ∞ 2
a20 X a2k
X
2πkτ bk 2πkτ
Rxx (τ ) = + cos + cos
4 2 T 2 T
k=1 k=1
∞
X 2πkτ
+ ak bk sin . (8.42)
T
k=1
The term on the last line of Eq. (8.42) originates from cross correlation of
sine and cosine terms of the series, which can be computed by integrating
ak bk sin(ω0 t) cos(ω0 (t + τ )) over one period.
500
a 400
b
5
300
200
0
100
-5 0
-100
-10
-200
0 2 4 6 8 0 10 20 30 40 50 60
300 7000
250
c 6000
1 d
200 5000 0
150 4000
-1
100 3000
0 2 4 6 8
2000 time (a.u.)
50
1000
0
0 2 4 6 8 0 10 20 30 40 50 60
time (a.u.) frequency (a.u.)
Figure 8.4: Detection of a noisy signal x(t) via autocorrelation. (a) Cosine wave
with frequency ν = 3 a.u. and unity amplitude superimposed by white noise with
S/N = (1/3)2 (simulation with 1024 data points). (b) Spectrum of the noisy signal
obtained by DFT. (c) Autocorrelation function Rx x(τ ) of the noisy signal. (d)
Spectrum of the autocorrelation function. The approximate frequency ν0 = 3.05
a.u. can be determined from the clearly visible peak. The inset shows the cross
correlation function Rzx (τ ) obtained with a subsidiary signal z(t) = cos(2πν0 t).
Note that phase and amplitude of Rzx (τ ) have noise-related deviations from the
corresponding parameters of x(t).
105
This property of the autocorrelation function can be used to detect a small
oscillatory signal superimposed by strong noise. For example, the oscilla-
tory signal with 1024 data points and a period 0.33 a.u. at signal-to-noise
ratio of (1/3)2 is not at all reccognizable by looking at Figure 8.4a. In the
spectrum obtained by DFT (Figure 8.4b), a peak at frequency 3 a.u. can be
recognized, but at this signal-to-noise ratio it is hard to be certain that this
is due to a true signal. Surprisingly, the oscillation is rather well seen in the
autocorrelation function of the signal (Figure 8.4c) and from the DFT of the
autocorrelation function (Figure 8.4d) the frequency can be extracted with
rather good precision. Cross correlation of the original signal with a sub-
sidiary function with this frequency provides an estimate for the oscillatory
function, which has approximately 20% amplitude error and 45◦ phase error
in this particular case. Larger errors can occur for this noise level (the errors
are themselves stochastic) and the frequency estimate can also have a larger
error, leading to beating in the cross correlation function Rzx . However, the
presence of the signal is always reliably detected.1 .
This can be understood by considering the total signal y(t) as a sum of the
wanted signal x(t) and noise n(t),
Since x(t) and n(t) are uncorrelated, the terms Rxn (τ ) and Rnx (τ ) vanish.
The term Rnn (τ ) consists of a large peak at τ = 0 (in the center of the plot
in Figure 8.4c), and much smaller noise at other times which stems from
finite-time observation. An estimate ν0 for the frequency of the oscillation
can be directly read off, determined by fitting, or taken from the spectrum
of the autocorrelation function.2 This estimate may be slightly erroneous
due to the noise in the DFT spectrum.
We can now construct a subsidiary signal z(t) = cos(2πν0 t). The cross
correlation function
1
Ryz (τ ) = x(τ ) (8.45)
T
1
This is no longer true for 1024 data points at S/N = (1/4)2
2
It is as well recognizable in the power spectrum.
106
is an estimate of the original signal x(t). Due to the error in ν0 and finite
observation time, amplitude and phase of Ryz (τ ) are slightly erroneous.
107
Chapter 9
Alternatives to
one-dimensional Fourier
transformation
108
9.1.2 2D spectroscopy
Data from some non-linear spectroscopic experiments can be processed by
two-dimensional FT, a concept that can be extended to multi-dimensional
FT. We consider a spectroscopic experiment where coherence on a transition
(m, n) evolves for time t1 with angular frequency ωmn and is then transferred
to a transition (p, q) where it evolves for time t2 with angular frequency ωpq
before being detected. Two types of coherence transfer allow for recognizing
the correlation between the two transitions from the dependence of the signal
on times t1 and t2 .
Phase modulation
In the first transfer type, the phase of the coherence on transition (p, q)
immediately after the transfer is given by
(0)
where φmn is the initial phase of the coherence on transition (m, n) at time
(0)
t1 = 0 and φpq is a constant phase shift introduced by the transfer. After
evolution, at detection time t2 , the phase of the coherence is given by
If detection causes a phase shift, we can include this by redefining φ0pq . The
signal eiφ2 can be written as
n o
hp (t1 , t2 ) = exp i ωmn t1 + φ(0)
mn + ω pq t2 + φ (0)
pq
n o n o
= exp i ωmn t1 + φ(0)
mn exp i ω pq t2 + φ (0)
pq
= c1 c2 − s1 s2 + ic1 s2 + is1 c2
= c1 (c2 + is2 ) + is1 (c2 + is2 ) . (9.3)
with
c1 = cos ωmn t1 + φ(0)
mn , s 1 = sin ωmn 1t + φ (0)
mn
c2 = cos ωpq t1 + φ(0)
pq , s 2 = sin ω t
pq 1 + φ (0)
pq . (9.4)
109
If required, coherence decay according to Eq. (1.14) can be accounted for
by multiplication with the decay factor
t1 t2
d(t1 , t2 ) = exp − exp − . (9.5)
T2,mn T2,pq
Amplitude modulation
In the second transfer type, the phase of the coherence on transition (p, q)
(0)
after the transfer is independent of t1 (φpq ), whereas the amplitude is given
by c1 . In this case, the signal without decay is given by
and coherence decay can again be accounted for by the factor defined in Eq.
(9.5).
2D Fourier transformation
110
then contains a single peak at (|ωmn |, ωpq ). Information on the sign of ωmn
is lost.
a b
c d
For phase modulation, the same peak is obtained, but according to the last
line of Eq. (9.3), we have to consider an additional term is1 (c2 + is2 ).
With respect to the t2 dimension, this term behaves the same way as the
other one. Hence, before FT along the t1 dimension we have an imaginary
part that is a sine modulated spectrum with a single peak at frequency
ωpq . A problem arises, since the first FT also generates an imaginary part,
which contains a dispersion peak at frequency ωpq . If Eq. (9.7) is strictly
adhered to, the product of this dispersion spectrum with is2 is a real input
for the FT along the t1 dimension. This leads to superposition of a negative
dispersion component D1 D2 with the wanted positive absorption component
A1 A2 , which is termed a phase-twisted lineshape (Figure 9.1). Dropping
the imaginary part after the first FT does not help in this case, since the
dispersion lineshape along t1 remains admixed. The problem can only be
circumvented if a mirror image h0p (t1 , t2 ) can be generated in which ωmn is
replaced by −ωmn . Sometimes this is possible by appropriate phase changes
of the excitation pulses. If so, the sum of phase-modulated signal and mirror
image correspond to an amplitude-modulated signal.
111
9.1.3 Deficiencies of spectrum estimators
Both one- and two-dimensional FT do not presuppose that the signal is a
linear combination of damped harmonics in time domain and thus a lin-
ear superposition of Lorentzian peaks in frequency domain. The numerical
versions DFT and 2D DFT provide estimates for the frequency response
(spectrum) of the system, whatever shape this response happens to have.
Such spectral estimators work equally well for spectra with a small, known
number of resolved transitions and spectra with a large unknown or even
infinite number of overlapping transitions.
The price paid for this universality is a fixed relation between relative fre-
quency resolution and the number of points acquired in time domain. The
sampling interval T is determined by the required bandwidth (Nyquist cri-
terion, Eq. (7.6)) and the frequency resolution by the number of data points
within the bandwidth, which is only twice as large as the number of data
points in time domain.1 This becomes a problem in multi-dimensional spec-
troscopy, where the size of the data array and the acquisition times for
indirect dimensions scale with the product of the number of time-domain
data points in all dimensions.2
A further disadvantage of FT arises from the assumption that the data are
known starting at time t = 0. Time-domain spectrometers tend to have a
certain deadtime that passes after high-power excitationn, before the small
response of the system can be detected. According to the shift theorem,
such a time shift leads to a phase shift. The phase shift is proportional
to frequency ωk throughout the whole lineshape of a peak centered at ωk .
Thus, it cannot be perfectly eliminated by a phase correction that is linear
with respect to the frequency variable ω. If broad peaks overlap, destructive
interference may occur. In that case, a small absorption peak superimposed
on the wing of the lineshape of a large absorption peak may appear as a hole
in the lineshape even in a magnitude spectrum (Figure 9.2a). The problem
can be understood as an admixture of dispersive cross terms Dm Dn to the
lineshape during computation of the magnitude spectrum, similar to the
admixture of a dispersive lineshape in 2D FT of phase-modulated signals.
In the one-dimensional case this admixture can be systematically averaged
1
Further zero filling does not improve resolution, but merely interpolates intermediate
data points in frequency domain.
2
This section and the following sections have profited from reading V.A. Mandelshtam,
Progr. Nucl. Magn. Reson. 38, 159-196 (2001) and from personal discussions with
Vladimir Mandelshtam.
112
by summing power spectra with different time delays and computing the
magnitude spectrum as the square root of the sum (Figure 9.2a). In practice,
the first power spectrum P0 (ω) is computed from the original signal with the
first data point at deadtime t0 = td . For each of the N − 1 subsequent power
spectra, with N being the number of original data points, a new time-domain
signal yk (t) is constructed by removing the first point of the current signal
and appending a zero. This corresponds to an initial time t0 = td + kT for
the k th data set. The cross term-averaged magnitude spectrum is obtained
as v
uN −1
uX
Mca =t Pk . (9.8)
k=0
a b
0 5 10 15 20 25 0 5 10 15 20 25
n (MHz) n (MHz)
N
X 1
y(t) = dk exp iωk − t , (9.9)
T2,k
k=1
113
where the number of lines N may be known or unknown. The total infor-
mation content in this signal are 3N real parameters, namely the amplitude
dk , angular frequencies ωk , and widths 2/T2,k of the N lines. If the sig-
nal is noiseless, 3N data points will theoretically suffice to determine the
parameters with arbitrary resolution, independently of the maximum obser-
vation time. This requires a signal processing algorithm that directly fits
the parameters rather than first computing a spectrum and determining the
lineshape parameters from this spectrum. Such an algorithm is a parameter
estimator.
Depending on the problem at hand, parameter estimators can suffer from
numerical instability in the presence of noise. In general, such instability is
encountered when the fit problem is non-linear. At first sight, the problem
defined by Eq. (9.9) appears to be non-linear. However, that problem can
be recast in linear form.
114
9.2.1 Linear prediction
We first consider a time-domain signal that arises from only a single term in
Eq. (9.9), sampled at a constant time interval T . This signal can be written
as
(k)
yl = Ae−lT /T2,k eiωk lT = AZkl , (9.10)
with
Zk = e(iωk −1/T2,k )T . (9.11)
The next point in the signal can be linearly predicted if we know Zk (note
(k)
that the yl are complex). For a noiseless signal we can compute Zk af-
(k) (k) (k) (k)
ter measuring two points, for instance y0 and y1 , since Zk = y1 /y0 .
Hence, if we know beforehand that the signal is a single damped harmonic
function and is noiseless, we can determine frequency and width with ar-
bitrary precision from two complex data points. The complex amplitude
(magnitude and initial phase φ0 ) is also encoded in the data points, but
does not influence Zk .
This property of damped harmonic functions can help to solve the deadtime
problem. Assume that we do not know data points y0 through y3 , but we do
(k) (k) (k) (k) (k)
know y4 and y5 . As Zk = y5 /y4 and, from Eq. (9.12), yl−1 = yl /Zk ,
we can iteratively predict all the missing points.
In practice, data are not noiseless, hence Zk is determined with a certain er-
ror. This leads to errors in the parameter estimates and to an extrapolation
error in linear prediction of missing parts of the signal that increases with
increasing time difference from the last known point. If we have more than
two points, we can reduce the error due to noise, since Zk can be computed
from any pair of points and in the mean of these values, noise will partially
average.
The concept carries over to the sum in Eq. (9.9), as the problem cast in this
form is linear. If we know N noiseless data points, we can predict earlier
and later data points, using
N
X
yl = bm yl−m , (9.13)
m=1
where the complex coefficients bm take the role previously taken by Zk and,
115
together with the known data points, encode the full information on ampli-
tudes, frequencies, widths, and phases of all N lines.4 In fact, the concept
of Eq. (9.13) can be applied to other problems than HIP, as long as they
are autoregressive.
As the signal is not noiseless, we shall need more data points to obtain a
reasonable quality of the linear prediction and reasonable estimates of the
line parameters. Hence, we define
M
X
yl = bm yl−m + el (M ≥ N ) , (9.14)
m=1
where the el account for the noise, which is not autoregressive. Note that the
el are not the noise amplitudes at point l themselves, but only the deviation
of the actual noise amplitude from a linear prediction from noise of the
preceding M points.
This equation can be restated in a matrix form appropriate for linear least-
square solution,
Tb = a − e , (9.15)
the index ranges arise from the condition that M should be as large as
possible and imply M = L/2, so that the length of vectors a and e is
also M . Since the rows of T are not necessarily linearly independent, the
equation cannot generally be solved by direct matrix inversion. In other
words, the system of equations may not have a unique solution. We can
ensure a unique solution by requiring that the sum of squares of the noise
4
The N complex coeffeicients bm encode 2N real informations, the same applies to the
N complex data points.
116
coefficients e attains a minimum,
where U (left singular vectors) and V (right singular vectors) are orthogonal
matrices, ∗ denotes the conjugate transpose, and S is a diagonal matrix of
singular values. SVD is related to the eigen problems of matrices TT∗
and T∗ T which provide the left and right singular vectors, respectively,
as eigenvectors. The non-zero eigenvalues in both eigenproblems are the
squares of the singular values.
With SVD, the solution of Eq. (9.17) is obtained as
117
3
Skk
0
4 8
k
12 16
a
c
b
d
0 1 2 3 4 5 -20 -10 0 10 20
time (a.u.) frequency (a.u.)
Figure 9.3: Linear prediction with singular value decomposition (simulation). The
inset is a plot of the singular values. (a) Original noisy time-domain data trace with
256 points. The real part is blue, the imaginary part red. (b) Points 17–256 were
predicted from coefficients bm that were computed from points 0–15 only. The sin-
gular value matrix S was truncated at N = 7. (c) Absorption spectrum computed
by DFT from the original data trace. (d) Absorption spectrum computed by DFT
from a data trace where points 16–255 were replaced by the linear prediction.
M
X
P (z) = z M
− bm z M −m , (9.21)
m=1
where b0 = 1. In computing a peak list, one can reject all complex roots
that correspond to exponentially rising rather than decaying signals, i.e. all
those one that have abs(Zk ) > 1.
The complex amplitudes ck eiφ0,k are obtained by least squares computation
from
M
X
yl = ck eiφ0,k zkl + wl , (9.22)
k=1
where the wl are the true noise contributions and ||wl ||2 is to be minimized.
However, this approach is sensitive to the ratio between sampling interval
and transverse relaxation times and, generally, producing a peak list in that
way from noisy data is not a very stable procedure.
If the signal contains noise, singular values beyond N approach a plateau
value larger than zero. The number of singular values that are significantly
larger than this plateau value is then an estimate for the number of lines
that have magnitudes significantly larger than the noise level. If the number
of lines N can be estimated in that way, signal-to-noise ratio of the predicted
118
points can be improved by setting all diagonal values Skk of S with k > N −1
to zero5 before computing the pseudo-inverse S+ and the bm from Eq. (9.19).
Such truncation of S does not reduce the number of non-zero prediction
coefficients bm .
Truncation does, however, reduce the noise in the predicted points. This
is illustrated in Figure 9.3. Linear prediction coefficients bm (m = 0 . . . 15)
were computed from a slightly noisy time-domain signal (a) corresponding
to the spectrum shown in (c). Matrix S was truncated by setting diagonal
elements Skk with k > 6 to zero. In this case, the original spectrum clearly
reveals that there are only seven lines. Points 16–255 of the original data
set were then replaced by the linear prediction according to Eq. (9.13). The
spectrum obtained by DFT from this data set (d) looks less noisy, although
the amplitudes, frequencies, and widths of the lines may not necessarily be
better defined.
119
interaction has the form
Z π/2
y(t)
3 cos2 θ − 1 ωdd t sin θdθ
= cos
y(0) 0
Z 1
3x2 − 1 ωdd t dx ,
= cos (9.23)
0
with
µ0 ~ 1 CAB
ωdd = γA γB 3 = 3 . (9.24)
4π r r
Since the permittivity of vacuum µ0 , Planck’s quantum of action ~, and the
gyromagnetic rations γA and γB are known constants, the only unknown is
the distance r. The spectrum Y (ω) that results by FT from Eq (9.23) is a
Pake pattern.
A distribution P (r) of distances r cannot easily be extracted from Y (ω). As
the time-domain signal depends only on r, it must be possible to transform
y(t)/y(0) to P (r). This corresponds to inversion of the integral equation
Z ∞ Z 1
y(t) 2
CAB
= P (r) cos 3x − 1 t dxdr . (9.25)
y(0) 0 0 r3
where we can make the identifications g(t) = y(t)/y(0), p(s) = P (r) and
where K(t, s) is the kernel. Inferring p(s) from g(t) is termed an inverse
problem.
In practical situations, g(t) is given as a vector g of discrete values. In
analogy to FT, the length of this vector and the sampling interval will have
some relation to the resolution and range of values in s domain. Hence, p(s)
is also well represented by a vector p, so that K(t, s) can be replaced by
a matrix. The discrete form of the Fredholm integral equation of the first
time thus reads
g = Kp . (9.27)
120
square matrix, which can be inverted.
121
9.4 Regularization
Regularization techniques stabilize the solution of ill-posed problems by in-
troducing additional information, in particular, additional constraints on
the solution. A typical restraints is smoothness of the solution, which can
be defined as a small square norm of the derivative or second derivative. If
the solution is known to be a probability density function p(s), the addi-
tional constraint p(s) ≥ 0 can be imposed, since probability densities cannot
be negative. A non-negativity constraint often stabilizes the solution very
strongly, since part of the problems with continuity arise from the fact that
negative contributions in s domain can partially offset positive contributions
at nearby points.
where D
b is a linear differential operator. Depending on the problem at
hand, D
b can be the first or second derivative. For the example of time-
domain signals from dipole-dipole interactions, the second derivative D
b =
d2 /dr2 performs best. The derivative or second derivative is a measure for
roughness, hence it minimization maximizes smoothness.
The regularization parameter α in Eq. (9.29) determines the relative weight-
ing between the square deviation of the simulated signal from the measured
signal (first term) and roughness (second term). In general, the optimum
choice of α depends on the expected error in g and the expected smooth-
ness of the data, for example, the width of the features that are expected
in the distance distribution P (r). Whereas the noise level in g can often be
estimated, the resolution of p(s) is usually not known beforehand.
122
-12
1
a a = 0.01 b
0.9 -16
log h
0.8
-20 aopt = 100
0.7
a = 1000
-24
0.6
0 0.5 1 1.5 2 -3 -2.8 -2.6 -2.4 -2.2
t (µs) log r
c d
P(r)
L curve method
b α ||2
η(α) = ||Dp (9.30)
123
Accordingly, on increasing α we expect a strong decrease in η, but only a
moderate increase in ρ. Vice versa, if α is too large, the roughness η will be
very small, but the square deviation between simulation and experimental
data ρ will be large. A further increase in α will then lead to only a moderate
decrease in η, but to a strong increase in ρ. A parametric plot of log η versus
log ρ will thus have an L shape, with the corner of the L corresponding
to an optimum regularization parameter αopt that balances errors due to
oversmoothing a errors due to overfitting the data by reproducing noise by
features in p(s).
124
Chapter 10
Spectrometers always perform some analog signal processing, since the de-
tected signal is transmitted and usually amplified before digitization and the
transmission lines and amplifiers have a certain bandwidth. Hence, some ef-
fective bandpass filter is invariably applied to the signal. In addition, it
may be necessary to apply a lowpass filter before analog-digital conversion
to avoid aliasing of high-frequency noise. Such aliasing would otherwise oc-
cur if the Nyquist frequency corresponding to the ADC sampling rate is
lower than the upper cutoff frequency of the effective bandpass filter.
Additional signal processing is more flexible and less expensive when it is
performed as digital signal processing. For instance, very high filter orders
can be realized digitally, while they would require a lot of electronic com-
ponents in analog realization. Digital signal processing is limited only by
computing times and this limitation is relevant only if the processing needs
to be performed in real time.
125
the same algorithms can be implemented on the digital signal processors
and field-programmable gate arrays introduced in the next two sections.
126
for typical microprocessors or digital signal processors. This flexibility does,
however, allow for more extensive parallelization than with DSPs and thus
for higher processing rates.
An example for an FGPA is the Virtex-5 processor of Xilinx, for which a
DSP development kit exists. This development kit interfaces to Matlab,
which can be used for filter design.
yn = S {xn } , (10.1)
where yn and xn denote the complete series. In the follwoing we shall use
the index k, as in xk to denote a single data point in the series.
We assume that the system is linear and time-invariant, and, unless noted
otherwise, causal. Most systems are also dynamic, i.e., their response de-
pends not only on the current input signal xk but also on previous input
values xl with l < k. We are most interested in stable systems. For a stable,
time-discrete system any bounded allowed time-discrete input xn results in
a bounded response yn .
Many of the concepts that we introduced for time-continuous LTI systems
can be easily extended to time-discrete LTI systems. For instance, we can
define a time discrete Dirac impulse as
(
1 for k = 0
δk = . (10.2)
0 for k 6= 0
The value yk of a series is thus the convolution of the series yn with a Dirac
127
impulse at time k,
∞
X
yk = yl δl−k = yn ∗ δn |n=k (10.3)
l=−∞
gn = S{δn } . (10.5)
yn = xn ∗ gn (10.6)
10.2.3 Causality
A time-discrete system is causal if and only if the impulse response vanishes
for negative indices,
gk = 0 for k < 0 (10.7)
10.2.4 Stability
A time-discrete system is stable if and only if the impulse response is abso-
lutely summable,
∞
X
|gk | < ∞ . (10.8)
k=−∞
128
10.2.5 Frequency response
If a time-discrete LTI system is excited by a harmonic oscillation
xk = Aeiω0 kT , (10.9)
129
10.1 for M = 2, where the boxes z −1 represent delay by the sampling interval
T , the triangles multiplication with the corresponding coefficient, and the
plus characters in circles addition (accumulation). Because of feeding output
signal back via the multipliers with ak coefficients, the filter is recursive.
This leads to the infinite impulse response.
A finite impulse response is thus obtained for n1 = n2 = 0 and m1 = 0, m2 =
M . Again, we can normalize all coefficients by a0 . The structure of such an
FIR filter is shown in Figure 10.2, where a slightly different representation
with Σ characters for accumulation is used.
10.3 z transformation
When discussing filters in Section 4.2.2, Laplace transformation was shortly
mentioned as a tool in filter theory. The transfer function of a filter was
written based on the poles and zeros in Eq. (4.44). For digital filters, we
consider this theory in more detail. The equivalent of Laplace transforma-
tion for time-discrete signals is the z transformation,
∞
X
Y (z) = Z {yn } = yk z −k . (10.11)
k=−∞
An acausal series yn can be split into a causal part y+ (k) with y+ (k) = 0
for k < 0 and an anticausal part y− (k) with y− (k) = 0 for k > 0. The
130
corresponding z transforms are
∞
X
Y+ (z) = yk z −k and
k=0
X0
Y− (z) = yk z −k (10.12)
k=−∞
whereas the anticausal part Y− exists inside a circle with center 0 and radius
r− = r (|z| < r− ) with
Hence, Y (z) exists in a circular ring between radii r− and r+ . We are mostly
concerned with strictly causal series, so that the z transform exists outside
a circle with radius r+ .
131
the argument ω can be considered to run from −Ω/2 to Ω/2 with Ω = 2π/T ,
so that ωT runs from −π to π.
Shift theorem
The time-shift expression has to be used with caution for causal series,
which would become acausal for l < 0. This is corrected by eliminating the
contribution from the acausal part
l−1
X
Z {yn+l } = z l Y+ (z) − yk z l−k for l ≥ 0 . (10.18)
k=0
Convolution theorem
The most important property relates to convolution, which was defined for
series in Eq. (10.4). We have
max {rx+ , ry+ } < |z| < min {rx− , ry− } . (10.20)
132
Correlation functions
where the z transform G(z) of the impulse response is the transfer function Systemfunktion
of the system. If the convergence region is also given, the transfer function
is a complete description of the system.
The transfer function can be easily obtained by z transformation of the
difference equation, Eq. (10.10). Using the shift theorem we obtain
n2
X m2
X
aν z −ν Y (z) = z −µ bµ X(z) , (10.26)
ν=n1 µ=m1
133
which can be inserted into Eq. (10.25) to give
Pm2 −µ
Y (z) µ=m1 bµ z
G(z) = = Pn2 −ν
. (10.27)
X(z) ν=n1 aν z
To simplify this equation, we choose the same index ranges for both sums,
namely M1 = min {n1 , m1 } and M2 = max {n2 , m2 }, with missing coeffi-
cients being set to zero. Hence,
PM2 −µ
µ=M1 bµ z
G(z) = PM2 . (10.28)
ν=M1 aν z −ν
Furthermore, we can shift all indices by −M1 (the system is time invariant)
and expand the fraction by z M with M = M2 − M1 to have non-negative
exponents of z. Thus, we obtain
PM M −µ
µ=0 bµ z P (z)
G(z) = PM = . (10.29)
ν=0 aν z M −ν Q(z)
M
X Aν
G(z) = A0 + , (10.31)
z − z∞ν
ν=1
134
Table C.3 in [Leon et al.]1 give
M
X
n−1
gn = A0 δn + Aν z∞ν u(n − 1) , (10.32)
ν=1
With Eq. (10.32) we can consider the stability criterion given in Eq. (10.8).
We find that a system is stable, if and only if the condition
∞
X
|z∞ν |n−1 < ∞ (10.34)
n=1
is fulfilled for all poles with indices ν. This leads to the stability criterion
In other words, all poles of the transfer function G(z) must be positioned
inside the unit circle for the system to be stable.
If the input signal is split to two systems and the outputs of these two
systems are additively combined (parallel systems), the transfer function is
the sum of the individual transfer functions
135
10.4.3 Frequency-response function, phase response, and group
delay
The continuous frequency response characteristic G∗ (ω) can be computed
by the z transform at z = iωT , as indicated in Section 10.3.2. This can
be seen by computing G∗ (ω) as the FT of the impulse response (see Eq.
(2.19)),
∞ ∞
X X −n
G∗ (ω) = gn e−iωnT = gn eiωT
n=−∞ n=−∞
iωT
= G(z = e ). (10.38)
This relation is applicable to any stable system, as in such a case the poles
are inside the unit circle, while the z = iωT describe a line on the unit circle.
Note also that this implies the limited frequency band of the discrete signal,
which we have already encountered in Section 7.1.3. For some discussions, it
is thus convenient to define a normalized angular frequency Ω = ωT , which
falls into an interval of width 2π.
The amplitude response M (ω), corresponding to the magnitude spectrum
of the time-cintinuous system, is given by
q
|G∗ (z)|2 = G(z)G∗ (z)
p p
M (ω) = = G(z)G(z −1 ) .
z=eiωT z=eiωT
(10.39)
The phase response is defined as
1 d d
τg = − φ(ν) = − φ(ω) . (10.41)
2π dν dω
If the group delay is not constant, different frequency components are de-
layed by the system for different times.
136
imaginary W=p/2
z01 P1
w0
z¥ 1 Q1
W0 W=0
Q2 1 real
P2
a2
z¥ 2
b2 e
z02
Figure 10.3: Pole-zero plot of a notch filter with two zeros on the unit circle at
Ω = ±π/2 and two zeros (1 − )e±iπ/2 . The value of the frequency response at
angular frequency ω0 , corresponding to normalized angular frequency Ω0 , can be
computed from the lengths P1 , P2 , Q1 , and Q2 of the zeros Z01 and z02 and poles
z∞1 and z∞2 by also taking into account angles α2 and β2 as well as the analogously
defined angles α1 and β1 .
and phase response can also be described in terms of the poles and zeros,
QM iωT − z
iωT µ=1 e 0µ
G∗ (ω) = G(z = e )= G 0 QM . (10.42)
iωT − z
ν=1 (e ∞ν )
In a pole-zero plot in the complex plane, the factors in the numerator and
denominator correspond to vectors between a zero or pole, respectively, and
the point on the unit circle that corresponds to current frequency ω0 . These
vectors are most conveniently described in polar coordinates, as illustrated
in Fig. 10.3. In such plots, a zero is marked by a circle, while a pole is
marked by a cross. The polar coordinates are given by
137
With these definitions we find
M
Y M
Y PM PM
βµ −i
iω0 T
Q−1 iγ+i αν
G z=e =C Pµ ν e
µ=1 ν=1 . (10.44)
µ=1 ν=1
M
Y M
Y
M (ω) = C Pµ Q−1
ν , (10.45)
µ=1 ν=1
M
X M
X
φ(ω) = γ + βµ − αν . (10.46)
µ=1 ν=1
The group delay τg is obtained as the negative derivative of Eq. (10.46) with
respect to ω0 .
Note than in a Bode diagram (see Section 4.2.1) of log M (ω) vs. ω the
components from the individual poles and zeros in Eq. (10.45) become
additive. A zero z0k on the unit circle leads to a factor Pk = 0 and thus
to complete suppression of the frequency corresponding to this point on the
unit circle. A pole on the unit circle leads to an infinite signal.
138
Based on these considerations, a symmetric notch filter can be designed
with two zeros at e±ΩA and two poles at (1 − )e±ΩA , as shown in Fig. 10.3
for ΩA = π/2. The notch frequency follows with the sampling interval T as
νA = ΩA /(2πT ). We now consider the transfer function of such a notch filter
with ΩA = π/2. In this case the zeros are z01 = i and z02 = −i, whereas the
poles are z∞1 = (1 − )i and z∞2 = −(1 − )i. Thus, the transfer function
is given by
(z − i) (z + i)
G(z) = . (10.47)
(z − (1 − )i) (z + (1 − )i)
In the complex plane, z01 has the coordinates (<{z01 }, ={z01 } = (0, 1)
whereas the point ω0 = Ω0 /T has the coordinates cos(Ω0 ), sin(Ω0 ). Hence,
q
P1 = (0 − cos(Ω0 ))2 + (1 − sin(Ω0 ))2
q
= cos2 (Ω0 ) + 1 − 2 sin(Ω0 ) + sin2 (Ω0 )
p p
= 2 − 2 sin(Ω0 ) = 2 − 2 cos(Ω0 − π/2) . (10.48)
P1 · P2
A(Ω) =
Q1 · Q2
p
2 − 2 cos(Ω0 − π/2)
= p
1 − 2(1 − ) cos(Ω0 − π/2) + (1 − )2
p
2 − 2 cos(Ω0 + π/2)
·p . (10.50)
1 − 2(1 − ) cos(Ω0 + π/2) + (1 − )2
139
broader stop bands. The two-pole notch filter as defined here has an even
amplitude characteristic and an odd phase characteristic. The strongest
phase shifts occur near the notch frequency.
1.4
1.2
0.8
A(W)
0.6
0.4
0.2
0
-p -p/2 0 p/2 p
W
140
noms P1 (z) and P2 (z), with
M1
Y
P1 (z) = (z − z0µ ) , |z0µ | ≤ 1 ,
µ=1
M
Y
P2 (z) = (z − z0µ ) , |z0µ | > 1 . (10.52)
µ=M1 +1
The polynom
M M
" !#
Y Y 1
P20 (z) = ∗ ∗
1 − zz0µ = −z0µ · z− ∗ (10.53)
z0µ
µ=M1 +1 µ=M1 +1
Here,
P1 (z)P20 (z)
GM = (10.55)
Q(z)
is the transfer function of a minimal phase system with the same amplitude
characteristic as the one of the original system. This follows from the unity
amplitude characteristic of the allpass and from fact that the mirrored roots
of P20 must all reside inside the unit circle, as the roots of P2 are all residing
outside the unit circle.
This principle can be used, for instance, to compensate for the amplitude
characteristic and construct a pure phase shifter. For that, we can compute
GM as shown above and invert it. The time-discrete system with transfer
function 1/GM is stable, since all its poles are inside the unit circle (they are
the zeros of the minimal phase system GM . By cascading the original system
with a system with transfer function 1/GM , the amplitude characteristic is
compensated and a phase shifter with the phase characteristic of the allpass
is obtained.
141
10.4.6 Transfer function of FIR systems
For an FIR sytem the current output value yn is determined by M + 1
earlier input values xn . . . xn−M , where we neglect the delay introduced by
processing. This is expressed by the difference equation
M
X
a0 yn = bµ xn−µ . (10.56)
µ=0
M
X
0
a0 z Y (z) = bµ z −µ X(z) , (10.57)
µ=0
which gives PM −µ
Y (z) µ=0 bµ z
G(z) = = . (10.58)
X(z) a0
We divide all coefficients bµ by a0 to obtain
M
X
G(z) = bµ z −µ . (10.59)
µ=0
142
phase response with respect to frequency and thus to a linear dependence
of the phase shift on frequency,
φ(ω) = cω , c ∈ R . (10.61)
which gives
φ(ω) = ∠G∗ (ω) = −ωT k , (10.64)
Hence,
φ(ω) = −ωT nn u − ωT α . (10.66)
It can be shown ([Leon et al.]) that a real-valued system can have linear
phase in sections only if
∞
X
gn sin [ωT (n − nν − α) + φ0 ] = 0 , (10.67)
n=−∞
where we have introduced an additive constant that allows for phase jumps
by 2π.
We assume that the impulse response of our filter is purely real (gn ∈ <)
and vanishes outside the interval [0, M T ]. Furthermore, we take M as even
and choose the normalized group delay k so that
M
nν = . (10.68)
2
143
From Eq. (10.65) it then follows that α = 0. With these assumptions we
can design linear-phase FIR filters with symmetric or antisymmetric impulse
response. We choose the case of symmetric impulse response,
M
gn = gM −n , n = 0, 1, . . . , . (10.69)
2
The condition for linear phase response, Eq. (10.67), thus becomes3
M/2
!
X M M
0 = gn sin ωT n − + gM −n sin ωT −n
2 2
n=0
M/2
X M
(gn − gM −n ) sin ωT n − . (10.70)
2
n=0
This condition is fulfilled indeed, since (gn − gM −n ) follows from Eq. (10.69).
We can also demonstrate the linear phase response by computing the fre-
quency response
M M
M
−iωT M
gn e−iωT (n− 2 )
X X
−iωT n
G∗ (ω) = gn e =e 2
n=0 n=0
M/2−1
M M M
gn e−iωT (n− 2 ) + gM −n e−iωT ( 2 −n)
X
= e−iωT 2 g M
2
n=0
M/2−1 h i
M M M
e−iωT (n− 2 ) + e−iωT (n− 2 )
X
= e−iωT 2 g M gn
2
n=0
M/2−1
M X M
= e−iωT 2 g M + 2 gn cos ωT n −
2 2
n=0
M
= e−iωT 2 Rg (ω) , (10.71)
144
10.5 Matlab tools for digital signal processing
10.5.1 Useful Matlab functions for system analysis and de-
sign
The following (selected) Matlab functions are convenient for numerical im-
plementation of the concepts treated in this Chapter
• poly finds polynom coefficients from zeros; given poles and zeros of the
system, the transfer function according to Eq. (10.59) or Eq. (10.60)
can be found
• filter filters a time series x and provides output y, when given coef-
ficients aν and bµ
• prony given the impulse response and the orders of the numerator and
denominator polynom, finds coefficients aν and bµ
• freqz returns the digital filter complex frequency response, when given
coefficients aν and bµ
• invfreqz returns real coefficients aν and bµ for a filter that has least
square deviation from a requested frequency response function, given
also the orders of numerator and denominator polynom (numbers of
poles and zeros); a discrete frequency axis must be specified
Of these functions, roots, poly, and filter are contained in the basic
Matlab program. The remaining functions require the Signal Processing
Toolbox.
145
10.5.2 Filter design with Matlab
The Signal processing toolbox in Matlab also contains tools with a graphical
user interface for filter analysis and design. The analysis and visualization
tool fvtool is convenient if one has a set of filter coefficients aν and bµ . Co-
efficients for Butterworth and Chebyshev type I and II filters mentioned in
Section 4.2.2 can be computed by functions butter, cheby1, and cheby2, re-
spectively. Cauer filters are also called elliptical filters and can be computed
by ellipt. Additional functions, such as buttord help in specification, by
returning the lowest order of a filter that fulfills certain specifications for the
stop and passband and for the edge frequencies of stopband and passband.
Digital filters are not restricted to the analog types discussed in Section 4.2.2.
It is possible to specify maximum deviations δp and δs for the passband and
stopband, such that
1 − δp ≤ Y ei Ω 1 + δp , 0 ≤ Ω ≤ Ωp
(10.72)
and
Y ei Ω
≤ δs , ΩS ≤ Ω ≤ π . (10.73)
146