KEMBAR78
SDFT Extend | PDF | Discrete Fourier Transform | Fast Fourier Transform
0% found this document useful (0 votes)
8 views24 pages

SDFT Extend

The document discusses digital filter designs for recursive frequency analysis, focusing on the discrete Fourier transform (DFT) and the stability and performance of various filter techniques. It reviews existing recursive methods, proposes enhancements, and introduces a novel technique for stabilizing infinite impulse response (IIR) DFT analyzers. The paper emphasizes the importance of numerical stability and presents simulation results to support its findings.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
8 views24 pages

SDFT Extend

The document discusses digital filter designs for recursive frequency analysis, focusing on the discrete Fourier transform (DFT) and the stability and performance of various filter techniques. It reviews existing recursive methods, proposes enhancements, and introduces a novel technique for stabilizing infinite impulse response (IIR) DFT analyzers. The paper emphasizes the importance of numerical stability and presents simulation results to support its findings.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 24

To appear in Journal of Circuits, Systems, and Computers (JCSC)

Accepted draft version, Aug. 2015


Available online at: http://arxiv.org/abs/1408.2294

DIGITAL FILTER DESIGNS FOR RECURSIVE FREQUENCY ANALYSIS

HUGH L. KENNEDY
Defence and Systems Institute, University of South Australia, Mawson Lakes
Adelaide, SA 5095, Australia
hugh.kennedy@unisa.edu.au

Received (12th Nov. 2014)


Revised (8th Aug. 2015)
Accepted (25th Aug. 2015)

Digital filters for recursively computing the discrete Fourier transform (DFT) and estimating the
frequency spectrum of sampled signals are examined, with an emphasis on magnitude-response
and numerical stability. In this tutorial-style treatment, existing recursive techniques are reviewed,
explained and compared within a coherent framework; some fresh insights are provided and new
enhancements/modifications are proposed. It is shown that the replacement of resonators by
(non-recursive) modulators in sliding DFT (SDFT) analyzers with either a finite impulse response
(FIR), or an infinite impulse response (IIR), does improve performance somewhat; however
stability is not guaranteed, as the cancellation of marginally stable poles by zeros is still involved.
The FIR deadbeat observer is shown to be more reliable than the SDFT methods, an IIR variant is
presented, and ways of fine-tuning its response are discussed. A novel technique for stabilizing IIR
SDFT analyzers with a fading memory, so that all poles are inside the unit circle, is also derived.
Slepian and sum-of-cosine windows are adapted to improve the frequency responses for the
various FIR and IIR DFT methods.
Keywords: digital filters, frequency estimation, FIR filters, Fourier transform, IIR filters, recursive
filters, stability.

1. Introduction
The remarkable efficiency of the fast Fourier transform (FFT) has confined alternative
implementations of the discrete Fourier transform (DFT), such as recursive filter banks
1. -4. , to just a handful of niche applications: where input signals are band limited (e.g.

in communications 5. -7. and radar/sonar systems 8. -10. ); where tight architectural


constraints are imposed (e.g. in mobile and embedded devices); and/or where low
input/output latency is required (e.g. in safety critical 11. , closed-loop control and
autonomous systems).
In many of the aforementioned digital systems, the DFT is used to form estimates of
internal or external system states, via a power density spectrum (PDS) estimate derived
from noisy sensor data. On the one hand, when an 𝑀-point recursive DFT is used to form
state estimates in a digital feedback-control system, the control loop and the sensor
signal digitizer (with an ideal low-pass anti-aliasing filter) are able to operate with a
common period of 𝑇 seconds. On the other hand, when an 𝑀-point batch FFT is used,
all 𝑀 samples must be collected and stored in a buffer before they are processed, thus
the controller must operate with a time period of 𝑀𝑇 seconds between each new
control command. This reduces the stability margins of the system, forcing a lower
loop-gain to be used, resulting in a ‘sluggish’ closed-loop system response. The use of
even relatively short temporal or spatiotemporal FFTs 12. , in guidance/autopilot
subsystems of fast-moving autonomous vehicles, where the primary sensor is a video
camera operating with a frame rate in the order of 100 Hz, may therefore be sufficient
to render the system useless. In many legacy systems however, data are transferred
from the sensor to the processor as blocks, rather than as sample-by-sample streams.
In these cases, latency is unavoidable and well suited to batch processing using an FFT.
Interest in recursive DFTs for coding, signal analysis and spectrum estimation
appears to have increased in recent years 1. -11. , motivated perhaps by current trends
in hardware technology where sampling frequencies of analog-to-digital converters
continue to increase, while clock frequencies of digital processors appear to be
plateauing after many years of near exponential growth. Sensors and processors are
also becoming tightly coupled and being developed as integrated systems, which
provides an opportunity for the introduction of new approaches to the management of
data flows. There is now, more than ever, a need to develop alternative DFT algorithms
that might be more efficient and easier to implement than the FFT on a new generation
of sensors and parallel computing platforms 8. ,10. .
In the next section some existing recursive DFT implementations are discussed and
compared in terms of their, (impulse- and frequency-) response characteristics and
their resistance to the accumulation of rounding errors. Despite the importance of
rounding error sensitivity in real systems 1. ,10. ,11. ,13. -15. ; this point has been
overlooked in many studies on recursive frequency analyzers 3. -5. ,16. -18. which are
highly susceptible to this problem due to the use of pole-zero cancellation on the unit
circle. This problem has been addressed for finite impulse response (FIR) Sliding DFTs
(SDFTs) in Refs. 13. and 19. through the use of modulators instead of resonators in
the filter bank (see Section 2.2). For the most part, this approach does improve
performance; however, the dc prototype filter used, still contains a pole on the unit
circle at 𝑧 = 1, which leaves the filters vulnerable to rounding errors, when the input
signal is corrupted by high-power impulsive noise, for example (see Section 3). It is
shown in this paper that this approach may also be applied to the infinite impulse
response (IIR) SDFTs (see Section 2.6).
Of the recursive DFT methods considered, the deadbeat-observer technique would
appear to be the most attractive 7. ,20. -23. , due to its low complexity and rounding error
immunity (see Section 2.3); however, it is not ideal for spectrum estimation because its
response is fixed and generally unfavorable due to the high side-lobe level of the
Dirichlet kernel 24. . Some novel ways of addressing this deficiency are discussed in this
paper (see Section 2.4), although freedom to configure and fine-tune this method, by
shifting its open-loop poles for instance, is complicated by the outer feedback loop. The
fading-memory IIR SDFT methods described in Refs. 16. -18. , do offer some tuning
flexibility; however they are susceptible to rounding error accumulation (see Section
2.6). The use of simple band-pass filter banks and exponentially windowed oscillators
is also discussed (see Section 2.5).
In addition to providing an introduction and overview of existing recursive
methods, with an emphasis on relative strengths, weaknesses and interrelationships, a
stabilized version of the fading memory technique described in Ref. 17. is presented
here (see Section 2.7). This new filter structure has a recursive frequency analysis stage
followed by a non-recursive mixing stage. The optimal mixing coefficients are
determined via a least-squares procedure in the time domain but applied as a
convolution in the frequency domain to ensure that all poles are inside the unit circle.
A somewhat less significant, but none-the-less useful, contribution and point of novelty
is the adaptation of Slepian windows to FIR frequency-sampling filter-banks (see
Section 2.2) and the use of sum-of-cosine windows to improve the response of the
various IIR SDFT methods such as Refs. 13. and 17. . To illustrate these main points,
results of computer simulations are presented and discussed in Section 3. The paper
closes with a summary, some recommendations, and concluding remarks, in Section 4.

2. Overview of Some Low-Latency DFT Methods

It is assumed in this paper that the DFT is used to construct a magnitude spectrum
estimate |𝑋̂(𝑛, 𝑘)|, of a one-dimensional signal; therefore filter characteristics that
promote the objective of detection and estimation in the frequency domain are favored
(e.g. low variance, low bias, reasonable main-lobe width and low side-lobe levels 22. ,24. )
even if they do not satisfy all the requirements of a formal time-to-frequency
transformation (e.g. orthogonality and perfect reconstruction 3. ,21. ). As the spectrum
is a function of both time and frequency, the methods are well suited to time-frequency
analysis 1. ; however, the focus here is mainly on steady-state performance for
approximately stationary signals. Scope is further limited to non-parametric and non-
adaptive methods, implemented using filter banks operating at a common sampling
rate. Other than the signal bandwidth and the time duration over which the signal is
approximately stationary, no prior information is used to design the filters. The filter
responses are time invariant and their maxima are uniformly spaced in frequency. All
methods have been factored to conform to a common conceptual architecture (see Fig.
1). This novel decomposition is intended to facilitate the analysis and design
comparisons described in this paper.
For notational and coding convenience, 𝑀 is assumed to be odd and all
coefficients/signals complex. Using complex variables, in mathematical equations and
source code, means that first-order analysis filters may be used; however, nearly twice
as many analyzers are required to cover the same frequency range. The associated
symmetry benefits are evident when the methods considered here are applied to multi-
dimensional signals 12. .
Throughout the remainder of this section, the Methods implemented in Section 3,
involving various permutations of new and existing techniques, are shown in bold
typeface. A simple numeric labeling scheme is used to avoid a proliferation of long
acronyms. The most basic analysis technique is described in the following subsection.
Section 2 is also used: to clarify the nomenclature presented in Fig. 1, to introduce the
function of some of the blocks in a familiar context, and to discuss the fundamentals of
non-parametric frequency analysis.
𝐻win 𝑋̂(𝑛, 𝑘)

𝑥(𝑛) 𝑥̂(𝑛 + 𝑙) 𝑥̂(𝑛)


+ 𝐻pre (𝑧) 𝐻ana (𝑧) 𝐻mix 𝑋̂raw 𝑐syn 𝑧 −𝑙

𝑥(𝑛) - 𝐾ℎ 𝑥̂(𝑛) 𝑥̂(𝑛)


𝑋̂(𝑛, 𝑘) 𝑐fbk 𝑧 𝑚syn

𝐾
𝑛: Discrete-time sample index; 𝑛 = 0 … ∞.
𝑘: Frequency ‘bin’ index.
𝐾: Maximum ‘measurable’ frequency bin index, −𝐾 ≤ 𝑘 ≤ +𝐾.
𝐵: Maximum frequency bin of ‘interest’ for a band-limited signal, −𝐵 ≤ 𝑘 ≤ +𝐵;
signal bandwidth = 2 𝐵 ⁄𝑀 (cycles per sample); 𝐵 ≤ 𝐾.
𝑀: ‘Nominal length’ (in samples) of the odd DFT; 𝑀 = 2𝐾 + 1.
𝑧 −1 : Unit delay.
𝑙: Prediction horizon (in samples); 𝑙 > 0.
𝐻(𝑧): Discrete-time transfer function.
𝑐: Gain factor.
𝑥(𝑛): Digitized input signal.
𝑥̂(𝑛): Estimate of the input signal.
𝑋̂raw (𝑛, 𝑘): Frequency spectrum estimate (raw).
𝑋̂(𝑛, 𝑘): Frequency spectrum estimate (windowed).
𝐻pre (𝑧): Pre-filter; a scalar normalizing gain factor in most cases.
𝐻ana (𝑧): A parallel bank of frequency-tuned analysis filters.
𝐻mix: Mixing matrix; only used in the stabilized IIR SDFT.
𝐻win : Window function; a convolution in the frequency domain.
𝑐syn : Synthesis factors; the ‘partial’ IDFT, vector input, scalar output.
𝑐fbk : Feedback gain; unity for FIR/IIR observer, zero otherwise.

Fig. 1. Generic DFT filter architecture.

2.1. FIR DFT


This non-recursive reference implementation (Method 1) is the simplest from a
conceptual perspective but the most complex from a computational perspective. The
time/frequency elements of the frequency-spectrum estimate 𝑋̂(𝑛, 𝑘), are scaled bins
of the DFT, computed via direct convolution in the time domain using a non-recursive
filter with a finite-impulse-response, i.e.
𝑋̂(𝑛, 𝑘) = ∑𝑀−1
𝑚=0 𝑤(𝑚)𝑏𝑘 (𝑚)𝑥(𝑛 − 𝑚) (1)
thus
𝐻ana,𝑘 (𝑧) = ∑𝑀−1
𝑚=0 𝑤(𝑚)𝑏𝑘 (𝑚)𝑧
−𝑚
(2)
where
𝑏𝑘 (𝑚) = 𝑒 𝑗2𝜋𝑚𝑘 ⁄𝑀 (3)
are the filter coefficients of the frequency analyzer with 𝑗 = √−1. The term ‘bin’ is used
here as a reminder that each filter responds not only to the frequency for which is
designed, but also to nearby frequencies.
The pre-filter for this technique simply normalizes the window response using
𝐻pre (𝑧) = 1⁄∑𝑀−1𝑚=0 𝑤(𝑚). If a rectangular window is applied in the time domain then
𝑤(𝑚) = 1 for 𝑚 = 0 … 𝑀 − 1 and 𝑤(𝑚) = 0 for 𝑚 = 𝑀 … ∞ thus 𝐻pre (𝑧) = 1⁄𝑀.
Evaluating 𝐻ana (𝑧) around the unit circle, i.e. substituting 𝑧 = 𝑒 𝑗𝜔 = 𝑒 𝑗2𝜋𝑓 into Eq. (2),
yields the frequency response of the 𝑘th frequency analyzer in the filter bank
sin{𝑀𝜋(𝑓−𝑘 ⁄𝑀 )}
𝐻ana,𝑘 (𝑓) = (4)
𝑀sin{𝜋(𝑓−𝑘 ⁄𝑀 )}

which is the Dirichlet kernel 𝒟𝑀 (𝑓) 24. , where 𝑓 is the normalized frequency (cycles
per sample) and 𝜔 is the angular frequency (radians per sample).
This response is undesirable because strong but distant signal components all
contribute to the output of the 𝑘th bin, due to the high side-lobes, which may mask the
presence of weak but nearby components. Use of a less ‘severe/abrupt’ or ‘tapered’
window function, such as the Slepian window (Method 2), improves the response by
lowering the side-lobe level at the expense of some main-lobe broadening. A broader
main-lobe makes it more difficult to resolve strong closely-spaced components, but it
also decreases the attenuation of components that are located midway between
adjacent bins 24. .
The finite impulse response of the Slepian window 𝑤(𝑚) for 𝑚 = 0 … ∞, is
constructed so that only the first 𝑀 coefficients are non-zero and so that the proportion
of its power density |𝑊(𝑓)|2 within the frequency band −𝑓∆ ≤ 𝑓 ≤ +𝑓∆ is maximized
24. . Convolving the frequency response of the Slepian window with the ideal response

of the 𝑘th frequency analyzer – a unit impulse 𝛿(𝑓 − 𝑘⁄𝑀) – shifts the window response
so that the power of the windowed analyzer is maximally concentrated within the band
𝑘
−𝑓∆ ≤ 𝑓 ≤ 𝑀𝑘 + 𝑓∆ . Note that this optimization criterion states nothing about the
𝑀
‘shape’ of the response in the pass band. The response in this band is therefore peaked,
not flat; whereas the response outside this band rolls off quickly. Both features are ideal
for frequency analysis. Slepian windows, or the first solution of the so-called discrete
(or digital) prolate spheroidal sequence (DPSS) were used in this study because the
maximal concentration criterion takes some of the guesswork out of the tuning process
25. ,26. . For completeness, the procedure for deriving Slepian windows is given in

Appendix A. In Ref. 27. , the design process has also been extended and generalized for
use in digital low-pass filter designs.
As the window is applied in the time domain, 𝐻win and 𝐻mix are simply the identity
matrices. The synthesis block applies a ‘partial’ inverse DFT (IDFT) operation, which
reconstructs the noise-free input signal at the (𝑛 + 𝑙)th sample, using only the
𝐵 𝐵
frequencies within the analysis band −𝑀 ≤ 𝑓 ≤ +𝑀 ; when the signal is known to be band
limited, 𝐵 < 𝐾. This method does not rely on the outer feedback loop to produce a
produce the PDS estimate; therefore 𝑐fbk = 0 and the synthesis operation is optional;
although it may be used if the system is also required to perform a low-pass filtering
function. However, the filter will have a very poor response – i.e. poor attenuation
outside the pass band and uneven gain with non-linear phase inside the pass band – if
a predictive filter is applied (𝑙 > 0). For best results, assuming a non-zero group delay
is tolerable, 𝑙 = −𝐾 should instead be used and the delay block omitted. This yields a
linear-phase FIR filter, with a reasonable magnitude response. The synthesis operation
is applied using
𝑥̂(𝑛 + 𝑙) = ∑+𝐵 ̂
𝑘=−𝐵 𝑐syn (𝑘)𝑋raw (𝑛, 𝑘) (5)

where
𝑐syn (𝑘) = 𝑒 −𝑗2𝜋𝑙𝑘 ⁄𝑀 (6)

and 𝑥̂ is an estimate of the noise-free signal 𝑥̃, see (D.3). Swapping 𝑋̂ for 𝑋̂raw may be
preferable in cases where tapered windows are not required (i.e. for improved
frequency resolution).

2.2. FIR SDFT


The sliding DFT (SDFT) 28. ,29. , sometimes referred to as the frequency-sampling
method in older literature 22. , uses a recursive comb pre-filter and a bank of recursive
first-order resonator filters with an infinite impulse response (IIR), to generate a bank
of frequency analyzers with a finite impulse response (FIR) and a Dirichlet-kernel
frequency-response. This FIR SDFT technique is implemented using
𝐻pre (𝑧) = 𝑀1 (1 − 𝑧 −𝑀 ) (7)
1
𝐻ana,𝑘 (𝑧) = 1+𝑎 −1 (8)
𝑘𝑧

where
𝑎𝑘 = −𝑒 𝑗𝜔𝑘 with 𝜔𝑘 = 2𝜋𝑘 ⁄𝑀 . (9)
The outer feedback loop is not utilized by this method.
In Eq. (8) the complex coefficient is in the feedback path of the resonator. This
means that the phase reference of the analyzer is always at the current sample (𝑚 = 0),
as per all of the other methods considered in this paper. This configuration is ideal for
filtering operations; however, if a fixed phase reference is preferred (e.g. at 𝑛 = 0) then
the complex coefficient should instead be placed in the forward path so that it appears
in the numerator and denominator of Eq. (8) 29. .
Unfortunately, the response of an FIR SDFT analyzer is generated using a
marginally-stable pole on the unit circle due to the resonator, cancelled by one of the
zeros created by the comb (Method 3). This leaves the analyzers vulnerable to the
accumulation of rounding errors if the zeros are unable to perfectly cancel the pole due
to finite machine precision. This does not mean that the filters are prone to ‘explosive’
instability; rather, a drift or bias may gradually appear over time. These errors are
immaterial if numeric variables are long and if processing intervals are short. The comb
pre-filter can be interpreted as being partially responsible for the errors – as it involves
the difference of two finite precision numbers. For signals containing only components
at the analysis frequencies – i.e. with 𝑀 being an integer multiple of all component
periods – the resulting difference should be zero when the system reaches steady-state,
however a non-zero value is output when finite precision is used. The following
resonators then integrate these errors, resulting in an offset.
It was shown in Ref. 13. that the stability of the FIR SDFT is improved if resonator
convolution – see Eq. (8) – is replaced by a modulator, an integrator (i.e. a pole at 𝑧 =
1), and a phase corrector (the FIR “mSDFT”). The modulating signal is a complex
sinusoid which may be generated recursively (Method 4) or pre-generated and stored
in a look-up table (Method 5). Around the same time, a similar approach was proposed
and used to transform the temporal dimension of a spatio-temporal filter 19. . The
modulating sinusoids are pre-generated for this filter; however, complexity is increased
through the use of a comb filter in each temporal frequency bin and a non-recursive
DFT in the spatial dimensions.
The frequency response of the various SDFT methods may be improved through the
application of sum-of-cosine windows in the frequency domain via a convolution
operation, which is applied by the 𝐻win block in Fig. 1 28. ,30. . These windows (e.g. Hann,
Hamming and Blackman) are designed to approximately cancel the side lobes of the
Dirichlet kernel using bins that surround the analysis bin (see Appendix C) 24. . However
Slepian-type windows may also applied if an ‘optimal’ approach is preferred; note that
mathematical optimality is not necessarily the same as system optimality, if low side-
lobes are more important than narrow main-lobes, for instance. A procedure for
deriving the 2𝐵win + 1 window coefficients, where 𝐵win ≤ 𝐵, is described in Appendix
B. This procedure produces the frequency-domain window coefficients 𝑤 ̃(𝑘) for
−𝐵win ≤ 𝑘 ≤ +𝐵win , that maximize the proportion of the power density in the
frequency band −𝑓∆ ≤ 𝑓 ≤ +𝑓∆ , given the constraints of a finite impulse response of
2𝐾 + 1 samples in the time domain and a window size of only 2𝐵win + 1 bins in the
frequency domain. Note that 𝐵win = 1 for the Hann and Hamming windows and 𝐵win =
2 for the Blackman window. Note also that if 𝐵 < 𝐾 then 𝐵win bins at the positive and
negative extremes of the spectrum cannot be windowed properly; rather than deriving
customized windows for these edge bins, they are left unprocessed here. This Slepian-
type window was applied to the FIR mSDFT, implemented using a pre-generated
modulator (Method 6).

2.3. FIR Deadbeat Observer


The deadbeat observer is a control-theoretic approach to the problem of frequency
analysis (Method 7) 7. ,20. -23. . A “deadbeat” response is produced when the output is
simply a delayed version of the input, at the sampling times at least. This is achieved by
placing all closed-loop poles at the origin of the 𝑧 plane. An “observer” is a state-space
control technique for estimating the internal states of a system, using a plant model
embodied within a filter which is driven by a prediction error. The immunity of this
method to rounding errors may also be interpreted using a fundamental control theory
principle, which asserts that the addition of an integrator to the forward path of a
control loop helps to drive steady-state tracking errors of the closed-loop system to
zero.
The deadbeat-observer approach to recursive DFT generation employs a bank of
integrators modulated by complex sinusoids, to generate a set of equally-spaced
marginally-stable poles around the unit circle for the open-loop system. A synthesis
operation is then performed to yield a one-sample prediction estimate, which is then
delayed by one sample, fed back and compared with the input. The bank of frequency
analyzers is then driven by the resulting error signal. When the closed-loop transfer
function of the 𝑘th frequency analyzer is determined 22. , it can be seen that the poles of
the other analyzers in the filter bank form a comb around the unit circle in the 𝑧 plane
(i.e. equally-spaced zeros), one of which cancels the 𝑘th pole, thus producing the
Dirichlet response in the frequency domain.
The observer methods are the only techniques that need the feedback loop, in all
other cases 𝑐fbk = 0. However; all methods may optionally use the time-averaged error-
signal as an indication of the spectrum quality – i.e. how well the specified frequency
components represent the input signal.

2.4. IIR Non-Deadbeat Observer


The situation described above is perfect for recursive DFT computation – with low
complexity and rounding error immunity; however for spectrum estimation, the
Dirichlet response is not always ideal. Many years ago when this method was first
proposed 20. , Bitmead was quick to point out that the finite-memory (i.e. FIR) deadbeat
response of this method may be sub-optimal and that a fading memory (i.e. IIR) might
be more desirable in some cases, where greater smoothing (in the time domain) was
needed to reduce the effects of random noise 31. . To deal with this problem he proposed
the use of a Kalman filter for frequency analysis 32. ; however, this approach is also not
well suited to spectrum estimation as it only reduces the variance of signal components
that coincide with the analysis frequencies (narrow main-lobes) and reduces the bias
due to the interference of other frequencies (low side-lobes). This approach is more
appropriate in applications, such as radar and communication systems, where the
component frequencies are known a priori.
The increased frequency selectivity of a slowly fading memory in a frequency analyzer
manifests itself as a narrower/sharper main-lobe with lower side-lobes 16. -18. ,32. .
Somewhat counter-intuitively perhaps, window functions in frequency estimation
operate by decreasing the frequency selectivity of the main lobe. As previously
mentioned, this is a side-effect of side-lobe reduction, but is not entirely undesirable as
it also increases the response to intra-bin components that might otherwise ‘slip
through the cracks’ and go undetected.
Peceli demonstrated that the response of the deadbeat observer could be improved
(low side-lobes without overly narrow main-lobes) by increasing the pole multiplicity
of the analyzers in the filter bank to second or third order 22. . This approach is clearly
very attractive; however, in this paper the focus is on banks of first-order IIR filters.
Returning now to Fig. 1: The deadbeat (Method 7) and non-deadbeat (Method 8)
observers are both implemented using: 𝐻pre (𝑧) = 1⁄𝑀; the same analysis filter as the
sliding DFT, see Eq. (8); 𝑙 = 1 in Eq. (5) and Eq. (6) and 𝑐fbk = 1. Other than increasing
the pole multiplicity as discussed above, the observer method is difficult to modify
because changes to the various transfer functions have unexpected and usually
undesirable consequences on the closed-loop transfer function. The poles must remain
on the unit circle, for the reasons discussed above which limits tuning flexibility;
however, two simple adjustments are proposed below to modify the response of the
deadbeat observer.
The first modification is the use of a band-limited bank of analyzers (i.e. using 𝐵 <
𝐾). From a DFT perspective, this involves the use of many samples to evaluate just a few
frequency bins. This approach is adopted for all of the other techniques considered in
this paper – primarily to reduce the computational load; however, for this particular
technique, the closed-loop response of each analyzer is also affected. In the 𝑧 plane, the
poles move radially outwards from the origin thus the system is no longer deadbeat (i.e.
FIR). In the time domain, the resulting response is less selective as it has a fading
memory, with an infinite impulse response (IIR); which increases the frequency
selectivity of the response. As discussed above, the lowering of the side-lobe levels is
particularly useful in frequency analysis applications; however, narrow main-lobes are
undesirable when the frequencies of signal components are unknown a priori. These
effects are most pronounced when 𝐾 ≫ 𝐵 where the poles approach the unit circle.
From a control-system perspective, this change in transient-response behavior in the
time domain may be interpreted as being due to the decreased gain applied by the pre-
filter in the forward path (𝐻pre (𝑧) = 1⁄𝑀). Unfortunately this band-limited approach
does result in some distortion of the main-lobe maxima (see Section 3).
As has already been mentioned, an overly ‘sharp’ main-lobe is not necessarily
desirable in frequency analysis. To moderate this effect, a second (complementary)
modification may be used. If the prediction horizon is increased beyond 1 sample (𝑙 >
1), then this decreases the frequency selectivity of the observer. As mentioned at the
end of Section 2.1, where the IDFT is discussed, when the synthesis sample moves away
from the center of the analysis window (where 𝑙 = −𝐾) the response deteriorates; thus
we have an alternative mechanism for conferring the main-lobe broadening effects of a
window function in an observer framework. If a frequency-domain window function is
also applied (see Appendix C), then this must be done outside the closed-loop circuit (as
shown in Fig. 1) 22. . The existence of the pre-filter gain and the prediction-horizon
parameters (see Fig. 1), is hidden in the standard representations of the deadbeat
observer used in the (historic and modern) literature on this topic; thus their possible
utility as response-tuning mechanisms has been largely overlooked.

2.5. IIR Band-Pass Filters


Another way to handle the problem of rounding error accumulation described in
Section 2.2 is to simply use a bank of first-order damped resonators, with poles inside
the unit circle, and to eliminate the comb pre-filter. The pole radius is selected to ensure
that there is reasonable overlap between the responses of adjacent filters, so that
frequency ‘coverage’ is reasonably uniform; a non-minimum-phase zero may also be
placed on the opposite side of the unit circle for rapid roll-off. Using a near-unity pole
radius 𝑟, yields highly frequency-selective filters. One complex ‘half’ of a real Goertzel
filter is produced when 𝑟 = 1; however, as a consequence of their ever-expanding
memory, Goertzel filters require the application of a window to concentrate their
impulse response in time 3. , using one of the approaches in Ref. 30. for instance,
although preferably one which does not involve poles on the unit circle 33. . However, if
the poles of the analyzers are inside the unit circle, then time concentration is a natural
consequence of the filter’s fading memory.
This approach is simply the recursive application of a one-sided exponential
window 𝑤(𝑚) = 𝑒 𝜎𝑚 (where 𝜎 < 0) to an un-damped resonator. When viewed in this
way, it is apparent that a variety of generalized windows with impulse responses of the
form 𝑤(𝑚) = 𝑚𝜅 𝑒 𝜎𝑚 , could be applied to customize the response, if the order of the
recursive window is increased using the ‘shape’ parameter 𝜅, with 𝜅 > 0 34. -36. .

2.6. IIR SDFT (Non-Stabilized)


Like the band-pass filters described in the previous section, the filters described here
have an exponentially fading memory; but unlike those analyzers, these analyzers are
orthogonal, which makes them more suitable for both spectrum sensing and transform
functions. They are also particularly useful in communications systems because
interference between equally-spaced channels is minimized. This technique features a
fading-memory pre-filter in series with a bank of (un-damped) resonators 17. . The comb
pre-filter has equally-spaced zeros on the unit circle with frequency-matched poles
inside the unit circle. The analyzers are tuned so that their poles cancel one of the zeros
on the unit circle, to form the main lobe of the frequency response. Unlike the Dirichlet
response of the FIR analyzers (see Sections 2.1, 2.2 & 2.3), the main-lobe of each IIR
analyzer is ‘sharpened’ by the pre-filter’s unmatched pole inside the unit circle. In
contrast to the deadbeat observer, the pole radius 𝑟 of the pre-filter is a directly
configurable parameter, allowing the time/frequency selectivity to be readily tuned.
The IIR SDFT response approaches the FIR response as 𝑟 approaches zero; furthermore,
the response of each analyzer is independent of 𝐵.
Like the FIR SDFT described in Section 2.2, this IIR SDFT technique also suffers from
rounding error accumulation due to pole-zero cancellation on the unit circle – a point
which is overlooked in Refs. 16. and 17. . Note that the method described in Ref. 16. is
essentially the same as that described in Ref. 17. ; however, each analyzer in Ref.
16. has a zero at 𝑧 = 1, which is apparently added to improve the phase response, and
poly-phase extensions are described in Ref. 17. .
This IIR SDFT with a fading-memory (Method 9), is implemented using,
(1−𝑟) 1−𝑧 −𝑀
𝐻pre (𝑧) = ∙ (10)
𝑀 1−𝑟𝑧 −𝑀

and
𝐻ana,𝑘 (𝑧) = 1+𝑎1 𝑧−1 (11)
𝑘

where
𝑎𝑘 = −𝑒 𝑗𝜔𝑘 and 𝑟 = 𝑒 𝜎𝑀 (12)
with 𝜎 being a forgetting-factor (𝜎 < 0), which is multiplied here by a factor of 𝑀 to
ensure that the pole radius is the same as the filters described in the next section. This
technique does not use the outer feedback loop (𝑐fbk = 0) thus the use of the synthesis
factors (𝑐syn ) is optional.
As described in Section 2.2 on the recursive DFT with a finite memory (the FIR SDFT,
Method 2), the analysis resonators in this recursive DFT with a fading memory (the IIR
SDFT, Method 9) may also be replaced by (pre-generated) modulators (the IIR mSDFT,
Method 10). It will be shown in Section 3 that this approach does improve stability
somewhat, for both response types, but not in all situations – A fully stabilized version
is presented in the next section. Note that in the recent literature the SDFT acronym is
usually used to refer to the FIR technique described in Section 2.2; however, in this
paper, due to their architectural similarity (different pre-filters), the less well-known
IIR technique described in this section (sometimes called the notch Fourier transform
16. ) is also referred to here as an SDFT variant.

Like the other DFT versions, application of a window is also optional; although doing
this (in the frequency domain) with just a few bins is counterproductive if the poles are
close to the unit circle because each main-lobe is very narrow and the side-lobes flat
and broad, with sharp notches at the frequencies of the other bins. As a consequence,
side-lobes do not cancel effectively as they do for Dirichlet kernels and the windowed
main-lobe features multiple maxima rather than a smooth monotonic decay down to
the edge of the first side-lobe. If sum-of-cosine windows are to be applied effectively
(see Appendix C), then the pole radius 𝑟, of the pre-filter should be set so that the
frequency response is at least somewhat similar to the Dirichlet kernel. A Hann window
was applied to the IIR mSDFT, implemented using a pre-generated modulator (Method
11).

2.7. Stabilized IIR SDFT


The problem of rounding-error accumulation that destabilizes the FIR- & IIR-SDFT
methods (Sections 2.2 & 2.6) is caused by the filter-bank integration of the rounding
errors generated by the comb pre-filter. The technique described here avoids this
problem by removing the damped comb pre-filter and replacing it with a frequency-
domain ‘mixing’ operation (Method 12). A fading-memory response is instead realized
using a bank of damped resonators. In effect, the mixing operation orthonormalizes a
bank of frequency-shifted ‘leaky’ integrators which act as band-pass filters. In this
regard, the technique is a combination of the techniques described in Sections 2.5 & 2.6
– yielding a stable orthogonal filter bank that is suitable for both estimation and
transformation functions.
The mixing operation is applied using a (2𝐵 + 1) × (2𝐵 + 1) ‘mixing matrix’ 𝐻mix ,
so that the 𝑘th windowed frequency bin is a linear combination of all 2𝐵 + 1 analyzer
outputs. Clearly, this operation is a significant overhead; however, the extra
computational effort is worthwhile in applications where: a one-sample latency is
required with bias-free long-term filter operation, if orthonormality is desirable and if
band-limited signals are expected (𝐵 ≪ 𝐾), to keep the computational complexity to a
minimum. The elements of the mixing matrix are derived in the time domain using a
weighted least-squares procedure that is described in Appendix D. The procedure fits a
basis set of complex sinusoids 𝜓𝑘 (𝑚) = 𝑒 𝑗𝜔𝑘 𝑚 , to the input signal 𝑥(𝑛), using an
exponentially decaying weighting function 𝑤(𝑚) = 𝑒 𝜎𝑚 = 𝑟 𝑚 . As shown in the
Appendix, the forgetting factor determines the pole radius (𝑟) of the analyzers in the
filter bank. As the pole radius approaches zero, errors may ‘creep’ into the mixing
matrix due to ill-conditioning, because too-many of the low-frequency basis functions
are too similar over the shorter time window. In the absence of numerical errors in the
design and filtering processes, the frequency responses of the stabilized and non-
stabilized versions of the IIR SDFT are effectively identical for |𝑓| < 𝐵 ⁄𝐾 (there are no
nulls outside this band for the stabilized version). The response of the stabilized IIR
SDFT is also somewhat similar to the response of the IIR non-deadbeat observer for all
𝑓.
This is the only technique that uses the mixing matrix 𝐻mix . For this technique
𝐻pre (𝑧) = 1, as 𝐻mix replaces the comb and also takes care of normalization. In the
filter bank,
𝐻ana,𝑘 (𝑧) = 1+𝑎1 𝑧−1 (13)
𝑘

where
𝑎𝑘 = −𝑒 𝜎+𝑗𝜔𝑘 = −𝑟𝑒 𝑗𝜔𝑘 . (14)
All other blocks are the same as those used in the previous section. Application of a
Hann window is optional.
Like all of the IIR methods discussed so far, this method yields very sharp main-
lobes as the poles approach the unit circle. This heightened frequency selectivity is ideal
for accurate phase and magnitude measurements in noisy environments when signal
frequencies are known a priori (e.g. active sonar, radar and communication systems).
However degraded sensitivity at non-design frequencies, relative to the Dirichlet
response of the FIR methods, means that signals at unknown frequencies may go
undetected.

3. Computer Simulations

3.1. Computer Code


Source code to evaluate the various algorithms, using the generic architecture shown in
Fig. 1, was built using a C++ compiler and executed on a PC with an i5-3570 CPU and a
64-bit operating system. Single-precision and double-precision versions were
instantiated to help reveal the relationship between numeric precision and bias due to
rounding error accumulation. The specified numeric type was used only in the filter
application and error analysis stages of execution; whereas, double precision was used
to generate the simulated input signal and design the digital filters for both the single
and double cases. No attempt was made to optimize the code at the instruction level, as
the emphasis was primarily on filter-response tune-ability and rounding-error
susceptibility. A non-negligible execution overhead was incurred through the use of
polymorphism, diagnostics, and non in-line functions, which were used to facilitate the
development process.

3.2. Method, Results and Discussion


Table 1 shows the dependence of the magnitude error |𝑋(𝑛, 𝑘)| − |𝑋̂(𝑛, 𝑘)|, of the
various DFT methods as a function of time-scale (𝑛) and noise environment, for a given
frequency bin. Data streams were processed continuously but divided into segments of
106 samples for simulation and analysis purposes. Two types of zero-mean additive
noise were examined: Gaussian-noise with a unity standard deviation, added to each
sample; and impulsive noise, uniformly distributed over ±1 × 106 , and added to the
first sample of each segment. The root-mean-squared error (RMSE) over each segment
was analyzed in the Gaussian case; the instantaneous error (Err.) was used in all other
cases. The input signal was generated using 𝐵 + 1 sinusoidal components with unity
magnitude, random phase and frequencies that coincided with each of the analysis bins
over 0 ≤ 𝑘 ≤ 𝐵. Real signals were used in all simulations. Over the timescale
investigated, the trends noted below for single-precision were barely noticeable for
double precision; therefore only single precision results are shown. Note that
windowed variants (Methods 2, 6 & 11) were not considered in this first numerical
experiment. The results in Table 1 reveal the following:
 Even in the absence of noise, Err. increases over time by two orders of magnitude
for Methods 3 & 9, due to rounding error accumulation.
 Methods 4, 5 & 10 do provide some resistance to error accumulation, in general;
however, their estimates are corrupted by impulsive noise. Methods 4 & 5 have
similar accuracy.
 Methods 8 & 12 show no signs of error accumulation and recover well after
impulse noise events. Method 12 is only slightly faster than Method 1; Method 8
is much faster than Method 1.
 In the Gaussian noise scenario, the IIR Methods 9, 10 & 12 are more accurate than
the FIR Methods 1, 3, 4 & 5, due to their narrow frequency response and long
impulse response – see Fig. 2 & Fig. 3. Fig. 2 also shows that there is some
asymmetric distortion of the main-lobe of Method 8.
In a second set of numerical experiments, the ability of the various DFT methods to
detect sinusoidal components of arbitrary frequency and magnitude was examined (see
Fig. 4). The input signal was the sum of two sinusoidal components with frequencies
midway between analysis bins: the first had unity magnitude and 𝑓 = 7.5⁄𝑀; the
second had a magnitude of 0.01 and 𝑓 = 17.5⁄𝑀. Only double precision was used. The
responses in Fig. 2 account for the trends shown in Fig. 4. The results in Fig. 4 suggest
the following:
 Tapered windows are required to detect weak components that are close to strong
components. In this respect, Methods 2, 6 & 11 are better than their un-windowed
counterparts: Methods 1, 5 & 10.
 The side-lobe reduction of Method 6 is nearly as good as Method 2.
 Method 11 has lower side-lobes and a broader main-lobe than Methods 2 & 6.
 Using a smaller 𝜎 for a larger 𝑟 in Methods 10-12 increases frequency selectivity
(at the expense of temporal selectivity) and decreases the response to arbitrary
signal components, i.e. components that are not near frequency bin centers.
 Methods 11 & 12 are indistinguishable, except for the last bin, due to numerical
errors in the design of the mixing matrix used in Method 12.
Note that using 𝐵 < 𝐾 transforms the FIR deadbeat observer into an IIR non-
deadbeat observer; thus Method 7 could not be examined in either of the numerical
experiments discussed above and only Method 8 results are presented.
Fig. 3 provides additional insight into the operation of the IIR SDFT methods. The
IIR SDFTs have an impulse-response with a ‘terrace’-like ‘envelope’, modulated by a
sinusoidal ‘carrier’. The stabilized and non-stabilized versions generate the same
envelope via different mechanisms (when 𝐵 = 𝐾). In the absence of windowing, the
magnitude over each step or block of length 𝑀, is constant and equal to
[(1 − 𝑟)⁄𝑀]exp(𝜎𝑀𝑛blk ) for 𝑛blk = 0 … ∞, where 𝑛blk is the block index. The quantity
2
|𝑋̂(𝑛, 𝑘)| produced by these IIR SDFTs may therefore be interpreted as being the 𝑘th
bin in a sliding exponentially-weighted Bartlett periodogram. Non-recursive versions
of this frequency estimator have been examined in Refs. 37. and 38. .

Table 1. Comparison of total execution time (s) and magnitude error of DFT techniques for 𝑘 = 16 with 𝐾 =
64 and 𝐵 = 32, in noise-free, Gaussian-noise and impulsive-noise scenarios.

RMSE Err. Err. Err. Exec.


Method
Gaussian a Impulsive b No Noise c No Noise d Time
1 1.80E-01 -3.60E-07 -3.60E-07 -1.20E-07 36
3 1.80E-01 -3.80E-02 -1.50E-02 -3.20E+00 5
4 7.80E-01 -7.80E-03 3.60E-07 6.00E-07 8
5 1.80E-01 -7.80E-03 7.20E-07 7.20E-07 5
8e 1.40E-01 -1.90E-06 -1.90E-06 -9.50E-07 5
9f 9.20E-02 -3.60E-02 -1.50E-02 -3.20E+00 9
10f 8.80E-02 -1.20E-03 7.20E-07 7.70E-07 5
12f 8.80E-02 -1.70E-06 -9.50E-07 -1.30E-06 28
a Averaged over 𝑛 = 106 to 𝑛 = 2 × 106 . b At 𝑛 = 2 × 106 . c At 𝑛 = 106 . d At 𝑛 = 108 . e With 𝑙 = 1. f With
𝜎 = 1⁄(2𝑀).

Fig. 2. Ideal frequency responses of DFT filters for 𝐾 = 8, 𝐵 = 4, and 𝑘 = 2. Inset shows detail of main-lobe.
Fig. 3. Impulse response of a non-stabilized IIR SDFT filter (Method 9), for 𝐾 = 64 (𝑀 = 129), 𝜎 = −1/𝑀,
and 𝑘 = 2. Application of the Hann window function via a convolution in the frequency domain (Method 11)
modulates each rectangular block of constant magnitude by the sum-of-cosines taper.

Fig. 4. Detection simulation for 𝐾 = 64 and 𝐵 = 32.


4. Conclusion and Summary

4.1. Impulse Response and Filter Structure


Of the DFT techniques considered here, Methods 1-7 have an FIR, whereas Methods
8-12 have an IIR; furthermore, Methods 1 & 2 are non-recursive, whereas Methods 3-
12 are recursive. Methods 7 & 8 employ an outer feedback loop. Methods 2, 6 & 11
employ tapered window functions. The various filter structures are summarized in
Table 2.

Table 2. Filter Structure Summary.

Outer
See Impulse Window
Method Class Recursive Feedback
Section Response Function
loop
Rectangular
1 DFT 2.1 FIR No Disabled
Time
Slepian
2 DFT 2.1 FIR No Disabled
Time
Rectangular
3 SDFT 2.2 FIR Yes Disabled
Time
Yes Rectangular
4 mSDFT 2.2 FIR Disabled
(Modulated) Time
Yes
Rectangular
5 mSDFT 2.2 FIR (Pre-Gen. Disabled
Time
Modulator)
Slepian
6 mSDFT 2.2 FIR Yes Disabled
Frequency
Deadbeat Rectangular
7 2.3 FIR Yes Enabled
Observer Time
Non-
Fading
8 Deadbeat 2.4 IIR Yes Enabled
Time
Observer
Fading
9 mSDFT 2.6 IIR Yes Disabled
Time
Yes
Fading
10 mSDFT 2.6 IIR (Pre-Gen. Disabled
Time
Modulator)
Yes Fading Time
11 mSDFT 2.6 IIR (Pre-Gen. & Hann Disabled
Modulator) Freq.
Yes
Fading
12 SDFT 2.7 IIR (Freq. Disabled
Time
Mixing)

4.2. Computational Complexity


The recursive DFT techniques investigated all have the potential to be faster than the
FFT for band-limited low-pass signals where filter-banks with 𝐵 ≪ 𝐾 may be used.
Their structure also makes them amenable to parallel implementation. With the
exception of the stabilized IIR SDFT (Method 12), which is somewhat slower than the
other techniques and has a complexity proportional to (2𝐵 + 1)2 , all of the recursive
techniques have complexity that is proportional to (2𝐵 + 1), in contrast to the non-
recursive FIR DFT (Method 1) which has complexity proportional to (2𝐵 + 1)𝑀. Even
when there is no execution-speed advantage to be gained, relative to the FFT which has
𝑀log 2 𝑀 complexity, recursive DFT techniques are more appropriate in closed-loop
control applications where low measurement latency is desirable.
Unlike all other FIR methods considered, the deadbeat observer technique (Method
7) cannot be accelerated by simply reducing the number of frequency bins evaluated,
due to the coupling between all analysers via the outer feedback loop; using 𝐵 < 𝐾 in
the filter bank ‘sharpens’ the frequency response and ‘smears’ the impulse response so
that it becomes infinite in duration (i.e. IIR), for each closed-loop analyser.

4.3. Numerical Stability


Computer simulations confirmed that the FIR and IIR SDFTs (Methods 3-6 and
Methods 9-11, respectively) are susceptible to rounding error accumulation when
single-precision floating-point arithmetic is used. Error accumulation is very gradual
for all methods when double precision is used; these techniques may therefore be used
with confidence to process short data sets, such as the pixels in a digital image or an
audio file. Replacement of resonators by modulators does improve stability somewhat;
however, high-power impulsive noise/interference, introduces errors which are never
‘forgotten’, due to the finite machine precision used in the running sums. This suggests
that the FIR and IIR SDFTs proposed so far in the literature are unsuitable in systems
intended to handle inputs with a high dynamic range. Even if these anomalous inputs
(perhaps due to ‘upstream’ system exceptions) are extremely unlikely events, their
eventual occurrence means that the filter bank, or the entire system, must be reset if it
is to continue functioning as designed.
The use of simple band-pass filter-banks with all poles inside the unit circle may be
an appropriate solution when only crude indications of channel occupancy are
required. However, the proposed stabilized IIR SDFT (Method 12), where the filter-
bank is orthonormalized in the frequency domain, is a better solution when a more
accurate/precise estimator or transformation is required. The outer feedback loop
ensures the long-term stability of the (FIR and IIR) observer in all cases examined
(Methods 7 & 8, respectively).

4.4. Magnitude Error


For a given frequency bin, using a recursive filter structure (Methods 3-12) allows a
longer average impulse response duration to be used (i.e. decreased temporal
selectivity), without incurring an extra computational cost, which reduces the expected
random error in the magnitude estimate due to random noise. Furthermore, the
availability of the pole radius parameter (𝑟 = 𝑒 𝜎 ) in the IIR techniques (Methods 9-12)
allows the time/frequency selectivity to be fine-tuned for a given combination of 𝐵 & 𝐾
paramters, i.e. for a particular distribution of frequency bins. Bias errors due to other
frequency components are also reduced, as the effective length of the impulse response
increases, due to the narrower main lobe and the lower side lobes. The application of a
tapered window function (e.g. Slepian or Hann), in either the time (Method 2) or
frequency domain (Methods 6 & 11), allows biases caused by the side lobes of other
‘distant’ frequency components to be further reduced, at the expense of a decreased
ability to resolve closely-spaced frequency components, due to a broadening of the
main-lobe. Of course, the aforementioned benefits are completely lost, however, if
rounding errors are permitted to accumulate in recursive implementations. Output
characteristics of the various filters are summarized in Table 3.

Table 3. Filter Output Summary.

Susceptible to Rounding
Method Main Lobe Side Lobes
Error Accumulation
1 Narrow High No
2 Wide Low No
3 Narrow High Yes
4 Narrow High Yes (In Some Cases)
5 Narrow High Yes (In Some Cases)
6 Wide Low Yes (In Some Cases)
7 Narrow High No
8 Sharp Flat No
9 Sharp Flat Yes
10 Sharp Flat Yes (In Some Cases)
Wide & Non-
11 Low Yes (In Some Cases)
Monotonic
12 Sharp Flat No

4.5. Frequency Response Flexibility


The frequency-domain Slepians were used to reduce the side-lobes of the Dirichlet
kernel (Method 6). This approach may be applied to the output of any of the recursive
techniques where the response is FIR (e.g. the FIR SDFT and the FIR deadbeat observer)
or approximately FIR (e.g. the IIR SDFT for small 𝑟). Sum-of-cosine windows (see
Method 11) may be applied in the usual way, and for the usual reasons, to the spectral
estimates output by all techniques; however, the benefits are limited in IIR cases when
the poles are close to the unit circle. The various IIR DFTs (Methods 8-11) may be
tuned to be very frequency selective (narrow main-lobes with low side-lobes) with
good noise rejection and are appropriate when the frequencies of the signal
components are known a priori, in communications and radar applications, for
example. In these cases, the bins may be placed to coincide with the expected
component frequencies 16. . Unlike the IIR SDFTs, it is not possible to change the
frequency response of the deadbeat observer (Method 7) by adjusting the pole
positions of the open-loop analyzers directly; however, the closed-loop poles may be
shifted indirectly, the impulse response changed from FIR to IIR, and the side-lobes of
the frequency response lowered, using 𝐵 < 𝐾; furthermore, the main-lobes are
broadened, for better detection of components with unknown frequencies, by
increasing the prediction horizon using 𝑙 > 1.
Appendix A. Derivation of the Time-Domain Slepian Window
The discrete-time transfer-function of a non-casual FIR window-function is
represented in the 𝒵 domain using
𝑊(𝑧) = 𝑧 𝐾 ∑𝑀−1
𝑚=0 𝑤(𝑚)𝑧
−𝑚
= ∑+𝐾
𝑚=−𝐾 𝑤(𝑚)𝑧
−𝑚
. (A.1)
Evaluating 𝑊(𝑧) around the unit circle, i.e. substituting 𝑧 = 𝑒 𝑗𝜔 = 𝑒 𝑗2𝜋𝑓 into Eq. (A.1),
yields the frequency response
𝑊(𝑓) = ∑+𝐾
𝑚=−𝐾 𝑤(𝑚)𝑒
−𝑗2𝜋𝑚𝑓
. (A.2)
The power density spectrum of this window is therefore
𝑃(𝑓) = |𝑊(𝑓)|2 = 𝑊 ∗ (𝑓)𝑊(𝑓) (A.3)
where the asterisk superscript denotes complex conjugation. Substituting Eq. (A.2) into
Eq. (A.3) yields
𝑃(𝑓) = ∑+𝐾 +𝐾 ∗
𝑚2 =−𝐾 ∑𝑚1 =−𝐾 𝑤 (𝑚2 )𝑒
+𝑗2𝜋𝑚2 𝑓
𝑤(𝑚1 )𝑒 −𝑗2𝜋𝑚1 𝑓
𝑗2𝜋(𝑚2 −𝑚1 )𝑓
= ∑+𝐾 +𝐾 ∗
𝑚2 =−𝐾 ∑𝑚1 =−𝐾 𝑤 (𝑚2 )𝑤(𝑚1 )𝑒 . (A.4)

The power over the interval −𝑓∆ ≤ 𝑓 ≤ +𝑓∆ is


+𝑓
𝑃∆𝑓 = ∫−𝑓 ∆ 𝑃(𝑓) 𝑑𝑓 (A.5)

which may be represented using vector/matrix notation as


𝑃∆𝑓 = 𝐰 † 𝐐∆𝑓 𝐰 (A.6)

where the dagger superscript denotes a Hermitian transpose and 𝐐∆𝑓 is a 𝑀 × 𝑀


matrix, with the element in the 𝑚2 th row and 𝑚1 th column being the integral
𝑚 ,𝑚1 +𝑓
𝑄∆𝑓2 = ∫−𝑓 ∆ 𝑒 𝑗2𝜋(𝑚2−𝑚1 )𝑓 𝑑𝑓

𝑒 𝑗2𝜋(𝑚2 −𝑚1 )𝑓∆ −𝑒 𝑗2𝜋(𝑚1 −𝑚2 )𝑓∆


= (A.7)
𝑗2𝜋(𝑚2 −𝑚1 )

which are real numbers. The Slepian window maximizes the power concentration
within the specified frequency interval. If the (band-limited) power over the interval
−𝑓∆ ≤ 𝑓 ≤ +𝑓∆ is 𝑃∆𝑓 and the (total) power over the interval −12 ≤ 𝑓 ≤ +12 is 𝑃1⁄2 then
their ratio 𝛼, with 0 ≤ 𝛼 ≤ 1, is the Rayleigh quotient
𝑃∆𝑓 𝐰 † 𝐐∆𝑓 𝐰
𝛼= = . (A.8)
𝑃 1⁄2 𝐰 † 𝐐1⁄2 𝐰

Due to the orthonormality of the complex sinusoids over the interval −12 ≤ 𝑓 ≤ +12,
𝐐1⁄2 = 𝐈, where 𝐈 is the identity matrix
𝐰 † 𝐐∆𝑓 𝐰
𝛼= . (A.9)
𝐰 †𝐰

The eigenvector corresponding to the greatest eigenvalue of 𝐐∆𝑓 maximizes this


quotient and it is the Slepian window function for 𝑚 = −𝐾 … + 𝐾. Finally, a delay of
𝐾samples is applied to yield a causal window for 𝑚 = 0 … 𝑀 − 1.
Appendix B. Derivation of the Frequency-Domain Slepian Window
Like the time-domain window, the frequency-domain window maximizes the
proportion of the power density response over the interval −𝑓∆ ≤ 𝑓 ≤ +𝑓∆ using a
window that is finite and non-zero in the time domain for 𝑚 = −𝐾 … + 𝐾; however, in
this case it is applied in the frequency domain using only 𝑘 = −𝐵win … + 𝐵win (see
Section 2.2). This additional constraint reduces the achievable concentration but good
responses are still possible, even for small 𝐵win . The window is derived using the
procedure described in Appendix A with a change of basis. The window is applied using
+𝐵win
𝑋̂(𝑛, 𝑘) = ∑𝑘́=−𝐵 ̃(𝑘́)𝑋̂raw (𝑛, 𝑘 − 𝑘́) .
𝑤 (B.1)
win

The corresponding time-domain representation of this window is


+𝐵
𝑤(𝑚) = ∑𝑘=−𝐵
win
win
̃(𝑘)𝑒 𝑗2𝜋𝑚𝑘 ⁄𝑀
𝑤 (B.2)

which is represented more compactly in vector/matrix form using


𝐰 = 𝐅𝐰
̃ (B.3)
where 𝐅 is a 𝑀 × (2𝐵win + 1) transformation matrix, with elements
𝐹(𝑚, 𝑘) = 𝑒 𝑗2𝜋𝑚𝑘 ⁄𝑀 . (B.4)
Substituting Eq. (B.3) for 𝐰 in Eq. (A.9) yields
̃ † 𝐅 † 𝐐∆𝑓 𝐅𝐰
𝐰 ̃
𝛼= ̃ † 𝐅 † 𝐅𝐰
̃
(B.5)
𝐰

which is solved as a generalized eigenvalue problem using 𝐅 † 𝐅 = 𝑀𝐈, due to the


orthogonal columns of the time-to-frequency transformation. As in the time-domain
case, the eigenvector corresponding to the greatest eigenvalue maximizes the
concentration in the pass band and it is the Slepian window function for 𝑚 = −𝐾 … + 𝐾
and 𝑘 = −𝐵win … + 𝐵win . The use of a non-causal window ensures that all elements of
𝐅 † 𝐐∆𝑓 𝐅 are real, which simplifies its eigen-decomposition. The window must be
delayed by 𝐾 samples in time domain to make it causal and so that its phase origin
coincides with that of the analyzers (at 𝑚 = 0). This is done via modulation with a
complex sinusoid in the frequency domain using a multiplier of 𝑒 −𝑗2𝜋𝑘𝐾 ⁄𝑀 for each bin
of the window.

Appendix C. Sum-of-Cosine Windows for Recursive DFTs


Sum-of-cosine windows are also applied using Eq. (B.1), where the elements 𝑤 ̃(𝑘) are
simply arbitrary constants 28. , with 𝑤 ̃(0) = 1 for proper normalization (0 dB gain at
dc); for example 𝑤̃(𝑘) = [12 1 12], for 𝑘 = −1 … + 1, for the Hann window. To shift the
phase origin to 𝑚 = 0 in the time domain, for the recursive filters considered in this
paper, a factor of 𝑒 −𝑗2𝜋𝑘𝐾 ⁄𝑀 is applied to the 𝑘th window coefficient. These windows
are not optimal in any sense, although they are useful to lower side-lobes none-the-less.
These windows benefit FIR and IIR DFT filters alike, provided the IIR poles are not too
close to the unit circle. For IIR filters, this ‘window function’ applies a periodic
modulation over an infinite extent in the time domain; however, the exponential decay
of the associated filter’s impulse response confines the window’s influence to recent
samples only.
Appendix D. Derivation of the Stabilized IIR SDFT
The ‘mixing matrix’ 𝐻mix is used in the frequency domain to orthonormalize a bank of
recursive fading-memory filters. Orthonormalization of the impulse and frequency
responses, ensures that

∑∞
𝑚=0 ℎana,𝑘2 (𝑚)ℎana,𝑘1 (𝑚) = 𝛿𝑘1 .𝑘2 (D.1)

and
𝐻ana,𝑘1 (𝑓) = 𝐻ana,𝑘1 (𝑘M2) = 𝛿𝑘1.𝑘2 (D.2)

where 𝛿 is the Kronecker delta. The elements of 𝐻win are derived here using (D.1) which
also ensures that Eq. (D.2) is satisfied.
Frequency analysis (and filtering) is interpreted here as a weighted least-squares
fitting problem. It is assumed that
𝑥̃(𝑛 − 𝑚) = ∑+𝐵 ∗
𝑘=−𝐵 𝛽𝑘 (𝑛)𝜓𝑘 (𝑚) (D.3a)
𝑥(𝑛) = 𝑥̃(𝑛) + 𝜀 (D.3b)
where 𝜓𝑘 (𝑚) = 𝑒 𝑗𝜔𝑘 𝑚 are the basis functions for the assumed sinusoidal signal model,
𝛽𝑘 (𝑛) are the model coefficients at the time of the 𝑛th sample, 𝑥̃(𝑛) is the true signal
value, 𝑥(𝑛) is the measured (noise-‘corrupted’) signal value, and 𝜀 is additive zero-mean
Gaussian noise 𝜀~𝒩(0, 𝜎𝑥2 ), with an unknown noise variance of 𝜎𝑥2 . Maximum
likelihood estimates 𝛽̂𝑘 (𝑛) of the parameters 𝛽𝑘 (𝑛), derived using all past and present
samples, are formed by minimizing the weighted sum-of-squared errors (SSE) where
SSE(𝑛) = ∑∞ ∗
𝑚=0 𝜖 (𝑛 − 𝑚)𝑤(𝑚)𝜖(𝑛 − 𝑚) (D.4)
and where 𝜖 is the residual of the least-squares fit
𝜖(𝑛 − 𝑚) = 𝑥(𝑛 − 𝑚) − 𝑥̂(𝑛 − 𝑚) (D.5)
and 𝑤(𝑚) is the error weighting function 𝑤(𝑚) = 𝑒 𝜎𝑚 , with the forgetting-factor
parameter 𝜎 < 0. In the context of our spectrum-estimation/signal-filtering problem, 𝜎
determines the pole radius (𝑟 = 𝑒 𝜎 ) of the filter bank in the 𝑧 plane, 𝜓𝑘 (𝑚) are the
frequency components that determine the pole angle (𝜔𝑘 = 2𝜋𝑘 ⁄𝑀) in the 𝑧 plane, and
𝛽𝑘 (𝑛) is the frequency spectrum coefficient of the 𝑘th bin at the 𝑛th sample 𝑋̂ (𝑛, 𝑘). As
the 𝜎, 𝐵 and 𝐾 parameters (where 𝑀 = 2𝐾 + 1 and 𝐵 ≤ 𝐾) determine the pole
positions of each first-order analyzer in the filter bank, this least-squares fitting process
may be viewed as being an optimal zero-placement technique. If high-order analyzers
(with non-unity pole multiplicity) are acceptable then 𝑤(𝑚) = 𝑚𝜅 𝑒 𝜎𝑚 with 𝜅 > 0 may
be used (see Section 2.5); however 𝜅 = 0 in what follows (to simplify the mathematical
working and the computer coding).
If the least-squares fit is performed over a finite time interval 𝑚 = 0 … 𝑀 − 1, then
the estimates of the model parameters 𝛽𝑘 (i.e. the spectrum coefficients) are
determined in the usual way using
̂ = 𝓑−𝟏 𝓐𝒙 .
𝜷 (D.6)
̂ is a column vector of length 2𝐵 + 1 containing the coefficient
In Eq. (D.6): 𝜷

estimates 𝜷 = [𝛽−𝐵 , … 𝛽̂𝑘2 , … 𝛽̂+𝐵 ] ; 𝒙 is a column vector of length 𝑀 containing
̂ ̂
the recent input signal history 𝒙 = [𝑥(𝑛), … 𝑥(𝑛 − 𝑚), … 𝑥(𝑛 − 𝑀 + 1)]† ; 𝓐 =
−1
𝚿 † 𝐖 ≡ 𝐻ana and 𝓑 = 𝚿 † 𝐖𝚿 ≡ 𝐻mix where 𝐖 is a square 𝑀 by 𝑀 matrix of zeros with
the weighting vector 𝒘 = [𝑤(0), … 𝑤(𝑚), … 𝑤(𝑀 − 1)] along its diagonal, and
𝚿 is an 𝑀 by 2𝐵 + 1 matrix with the element in the 𝑚th row and 𝑘1 th column equal to
𝜓𝑘∗ 1 (𝑚).
If the least-squares fit is now performed over an infinite time interval 𝑚 = 0 … ∞,
then the model parameters 𝛽𝑘2 are also estimated using (D.6) however the finite
summations in 𝓑 are now replaced by the infinite summations

ℬ𝑘2.𝑘1 = ∑∞
𝑚=0 𝜓𝑘2 (𝑚)𝑤(𝑚)𝜓𝑘1 (𝑚) (D.7)

which may conveniently be evaluated in the 𝑧 domain using


ℬ𝑘2.𝑘1 = 𝒵{𝜓𝑘2 (𝑚)𝑤(𝑚)𝜓𝑘∗ 1 (𝑚)}|
𝑧=1

𝜎+𝑗(𝜔𝑘2 −𝜔𝑘1 )
= 1⁄[1 − 𝑒 ]. (D.8)

And taking the 𝒵 transform of 𝓐 yields the bank of analysis filters with the 𝑘1 th element
being
𝒜𝑘1 = 𝒵{𝑤(𝑚)𝜓𝑘1 (𝑚)} = 1⁄[1 + 𝑎𝑧 −1 ] (D.9)

where 𝑎 = −𝑒 𝜎+𝑗𝜔𝑘1 .
The mixing matrix and a Hann-like window may be applied in a single operation
using 𝐻mix&win = 𝐻win 𝐻mix , where 𝐻win is a tri-diagonal matrix (for 𝐵 < 𝐾) with the
Hann window coefficients running along its diagonals (see Appendix C).

References
1. S. Assous and L. Linnett, “High resolution time delay estimation using sliding discrete Fourier
transform,” Digital Signal Process., vol. 22, no. 5, pp. 820-827, Sep. 2012.
2. A. van der Byl and M.R. Inggs, “Recursive sliding discrete Fourier transform with
oversampled data,” Digital Signal Process., vol. 25, pp. 275-279, Feb. 2014.
3. J. Alhava and M. Renfors, “Recursive Algorithms for Modulated and Extended Lapped
Transforms,” IEEE Trans. Circuits Syst. I, Reg. Papers, vol. 61, no. 1, pp. 191-201, Jan. 2014.
4. Chun-Su Park and Sung-Jea Ko, “The Hopping Discrete Fourier Transform [sp Tips&Tricks],”
IEEE Signal Process. Mag., vol. 31, no. 2, pp. 135-139, Mar. 2014.
5. S.-C. Lai, Wen-Ho Juang, Chia-Lin Chang, Chen-Chieh Lin, Ching-Hsing Luo and S.-F. Lei, “Low-
Computation-Cycle, Power-Efficient, and Reconfigurable Design of Recursive DFT for
Portable Digital Radio Mondiale Receiver,” IEEE Trans. Circuits Syst. II, Exp. Briefs, vol. 57, no.
8, pp. 647-651, Aug. 2010.
6. S.-C. Lai, S.-F. Lei, Wen-Ho Juang and Ching-Hsing Luo, “A Low-Cost, Low-Complexity, and
Memory-Free Architecture of Novel Recursive DFT and IDFT Algorithms for DTMF
Application,” IEEE Trans. Circuits Syst. II, Exp. Briefs, vol. 57, no. 9, pp. 711-715, Sep. 2010.
7. L. Varga, Z. Kollar and P. Horvath, “Recursive Discrete Fourier Transform based SMT
receivers for cognitive radio applications,” in Proc. Int. Conf. on Syst., Signals and Image
Process., pp. 130-133, 11-13 Apr. 2012.
8. M. Inggs, A. van der Byl and C. Tong, “Commensal radar: Range-Doppler processing using a
recursive DFT,” in Proc. Int. Conf. on Radar, pp. 292-297, 9-12 Sep. 2013.
9. S. Assous and L. Linnett, “High resolution time delay estimation using sliding discrete Fourier
transform,” Digital Signal Process., vol. 22, no. 5, pp. 820-827, Sep. 2012.
10. A. van der Byl, R.H. Wilkinson and M.R. Inggs, “Recursive Fourier Transform hardware,” in
Proc. IEEE Radar Conference, pp. 746-750, 23-27 May 2011.
11. H.A. Darwish and M. Fikri, “Practical Considerations for Recursive DFT Implementation in
Numerical Relays,” IEEE Trans. Power Del., vol. 22, no. 1, pp. 42-49, Jan. 2007.
12. T. Schwerdtfeger, S. Schauland, J. Velten, A. Kummert, “On multidimensional velocity filter
banks for video-based motion analysis of world-coordinate objects,” in Proc. Int. Workshop
on Multidimensional Syst., pp. 1-5, 5-7 Sep. 2011.
13. K. Duda, “Accurate, Guaranteed Stable, Sliding Discrete Fourier Transform [DSP Tips &
Tricks],” IEEE Signal Process. Mag., vol. 27, no. 6, pp. 124-127, Nov. 2010.
14. M. Tasche and H. Zeuner, “Roundoff Error Analysis of the Recursive Moving Window Discrete
Fourier Transform,” Advances in Computational Mathematics, vol. 18, no. 1, pp. 65-78, Jan.
2003.
15. Yisheng Zhu, Hong Zhou, Hong Gu and Zhizhong Wang, “Fixed-point error analysis and an
efficient array processor design of two-dimensional sliding DFT,” Signal Process., vol. 73, no.
3, Jan. 1999.
16. M.T. Kilani and J.F. Chicharo, “A constrained notch Fourier transform,” IEEE Trans. Signal
Process., vol. 43, no. 9, pp. 2058-2067, Sep. 1995.
17. A.R. Varkonyi-Koczy, “Efficient polyphase DFT filter banks with fading memory,” IEEE Trans.
Circuits Syst. II, Analog Digit. Signal Process., vol. 44, no. 8, pp. 670-673, Aug. 1997.
18. M.B. Malik, M.U. Hakeem, I. Ghazi and A. Hassan, “Recursive Least Squares Spectrum
Estimation,” in Proc. IEEE Int. Symp. on Ind. Electron., pp. 599-602, 9-13 Jul. 2006.
19. H.L. Kennedy, “Efficient velocity-filter implementations for dim target detection,” IEEE
Trans. Aerosp. Electron. Syst., vol. 47, no. 4, pp. 2991-2999, Oct. 2011.
20. G.H. Hostetter, “Recursive discrete Fourier transformation,” IEEE Trans. Acoust., Speech,
Signal Process., vol. 28, no. 2, pp. 184-190, Apr. 1980.
21. G. Peceli, “A common structure for recursive discrete transforms,” IEEE Trans. Circuits Syst.,
vol. 33, no. 10, pp. 1035-1036, Oct. 1986.
22. G. Peceli and G. Simon, “Generalization of the frequency sampling method,” in Proc. Instrum.
and Meas. Tech. Conf. IEEE, pp. 339-343, 1996.
23. A.R. Varkonyi-Koczy, G. Simon, L. Sujbert, and M. Fek, “A fast filter-bank for adaptive Fourier
analysis,” IEEE Trans. Instrum. Meas., vol. 47, no. 5, pp. 1124-1128, Oct. 1998.
24. F.J. Harris, “On the use of windows for harmonic analysis with the discrete Fourier
transform,” Proc. IEEE, vol. 66, no. 1, pp. 51-83, Jan. 1978.
25. D. Slepian, “Prolate spheroidal wave functions, fourier analysis, and uncertainty — V: the
discrete case,” The Bell System Technical Journal, vol. 57, no. 5, pp. 1371-1430, May-Jun. 1978.
26. T. Verma, S. Bilbao and Teresa H.‐Y. Meng, “The digital prolate spheroidal window,” in Proc.
IEEE Intern. Conf. on Acoustics, Speech, and Signal Processing, vol. 3, pp. 1351-1354, 7-10 May
1996.
27. A. Tkacenko, P.P. Vaidyanathan and T.Q. Nguyen, “On the eigenfilter design method and its
applications: a tutorial,” IEEE Transactions on Circuits and Systems II: Analog and Digital
Signal Processing, vol. 50, no. 9, pp. 497-517, Sep. 2003.
28. E. Jacobsen and R. Lyons, “The sliding DFT,” IEEE Signal Process. Mag., vol. 20, no. 2, pp. 74-
80, Mar. 2003.
29. E. Jacobsen and R. Lyons, “An update to the sliding DFT,” IEEE Signal Process. Mag., vol. 21,
no. 1, pp. 110-111, Jan. 2004.
30. B.G. Sherlock and D.M. Monro, “Moving discrete Fourier transform,” IEE Proc. Radar and
Signal Process. F., vol. 139, no. 4, pp. 279-282, Aug. 1992.
31. R.R. Bitmead, “On recursive discrete Fourier transformation,” IEEE Trans. Acoust., Speech,
Signal Process., vol. 30, no. 2, pp. 319-322, Apr. 1982.
32. R.R. Bitmead, Ah Chung Tsoi, and P.J. Parker, “A Kalman filtering approach to short-time
Fourier analysis,” IEEE Trans. Acoust., Speech, Signal Process., vol. 34, no. 6, pp. 1493-1501,
Dec. 1986.
33. J. Gronczyński, “Recursive Fourier transform algorithms with integrated windowing,” Signal
Process., vol. 87, no. 5, pp. 1003-1013, May 2007.
34. W. Chen, N. Kehtarnavaz and T.W. Spencer, “An efficient recursive algorithm for time-varying
Fourier transform,” IEEE Trans. Signal Process., vol. 41, no. 7, pp. 2488-2490, Jul. 1993.
35. S. Tomazic, “On short-time Fourier transform with single-sided exponential window,” Signal
Process., vol. 55, no. 2, pp. 141-148, Dec. 1996.
36. M.G. Amin and K.D. Feng, “Short-time Fourier transforms using cascade filter structures,”
IEEE Trans. Circuits Syst. II, Analog Digit. Signal Process., vol. 42, no. 10, pp. 631-641, Oct.
1995.
37. T.T.J.M. Peeters and O. Ciftcioglu, “Statistics on exponential averaging of periodograms,” IEEE
Trans. Signal Process., vol. 43, no. 7, pp. 1631-1636, Jul. 1995.
38. F.K. Tuffner, J.W. Pierre and R.F. Kubichek, “Computationally efficient updating of a weighted
Welch periodogram for nonstationary signals,” in Proc. Midwest Symp. on Circuits and Syst.,
pp. 799-802, 10-13 Aug. 2008.

You might also like