Least-squares spectral analysis

Last updated

The result of fitting a set of data points with a quadratic function Linear least squares2.svg
The result of fitting a set of data points with a quadratic function

Least-squares spectral analysis (LSSA) is a method of estimating a frequency spectrum based on a least-squares fit of sinusoids to data samples, similar to Fourier analysis. [1] [2] Fourier analysis, the most used spectral method in science, generally boosts long-periodic noise in the long and gapped records; LSSA mitigates such problems. [3] Unlike in Fourier analysis, data need not be equally spaced to use LSSA.

Contents

Developed in 1969 [4] and 1971, [5] LSSA is also known as the Vaníček method and the Gauss-Vaniček method after Petr Vaníček, [6] [7] and as the Lomb method [3] or the Lomb–Scargle periodogram, [2] [8] based on the simplifications first by Nicholas R. Lomb [9] and then by Jeffrey D. Scargle. [10]

Historical background

The close connections between Fourier analysis, the periodogram, and the least-squares fitting of sinusoids have been known for a long time. [11] However, most developments are restricted to complete data sets of equally spaced samples. In 1963, Freek J. M. Barning of Mathematisch Centrum, Amsterdam, handled unequally spaced data by similar techniques, [12] including both a periodogram analysis equivalent to what nowadays is called the Lomb method and least-squares fitting of selected frequencies of sinusoids determined from such periodograms — and connected by a procedure known today as the matching pursuit with post-back fitting [13] or the orthogonal matching pursuit. [14]

Petr Vaníček, a Canadian geophysicist and geodesist of the University of New Brunswick, proposed in 1969 also the matching-pursuit approach for equally and unequally spaced data, which he called "successive spectral analysis" and the result a "least-squares periodogram". [4] He generalized this method to account for any systematic components beyond a simple mean, such as a "predicted linear (quadratic, exponential, ...) secular trend of unknown magnitude", and applied it to a variety of samples, in 1971. [5]

Vaníček's strictly least-squares method was then simplified in 1976 by Nicholas R. Lomb of the University of Sydney, who pointed out its close connection to periodogram analysis. [9] Subsequently, the definition of a periodogram of unequally spaced data was modified and analyzed by Jeffrey D. Scargle of NASA Ames Research Center, [10] who showed that, with minor changes, it becomes identical to Lomb's least-squares formula for fitting individual sinusoid frequencies.

Scargle states that his paper "does not introduce a new detection technique, but instead studies the reliability and efficiency of detection with the most commonly used technique, the periodogram, in the case where the observation times are unevenly spaced," and further points out regarding least-squares fitting of sinusoids compared to periodogram analysis, that his paper "establishes, apparently for the first time, that (with the proposed modifications) these two methods are exactly equivalent." [10]

Press [3] summarizes the development this way:

A completely different method of spectral analysis for unevenly sampled data, one that mitigates these difficulties and has some other very desirable properties, was developed by Lomb, based in part on earlier work by Barning and Vanicek, and additionally elaborated by Scargle.

In 1989, Michael J. Korenberg of Queen's University in Kingston, Ontario, developed the "fast orthogonal search" method of more quickly finding a near-optimal decomposition of spectra or other problems, [15] similar to the technique that later became known as the orthogonal matching pursuit.

Development of LSSA and variants

The Vaníček method

In linear regression, the observations (red) are assumed to be the result of random deviations (green) from an underlying relationship (blue) between a dependent variable (y) and an independent variable (x). Then in a normed fitting, such as by the criterion of least squares, the data points (red) are represented by the line of normatively best fit (blue), from which there always remain "residuals" (green). Linear least squares example2.svg
In linear regression, the observations (red) are assumed to be the result of random deviations (green) from an underlying relationship (blue) between a dependent variable (y) and an independent variable (x). Then in a normed fitting, such as by the criterion of least squares, the data points (red) are represented by the line of normatively best fit (blue), from which there always remain "residuals" (green).

In the Vaníček method, a discrete data set is approximated by a weighted sum of sinusoids of progressively determined frequencies using a standard linear regression or least-squares fit. [16] The frequencies are chosen using a method similar to Barning's, but going further in optimizing the choice of each successive new frequency by picking the frequency that minimizes the residual after least-squares fitting (equivalent to the fitting technique now known as matching pursuit with pre-backfitting [13] ). The number of sinusoids must be less than or equal to the number of data samples (counting sines and cosines of the same frequency as separate sinusoids).

A data vector Φ is represented as a weighted sum of sinusoidal basis functions, tabulated in a matrix A by evaluating each function at the sample times, with weight vector x:

,

where the weights vector x is chosen to minimize the sum of squared errors in approximating Φ. The solution for x is closed-form, using standard linear regression: [17]

Here the matrix A can be based on any set of functions mutually independent (not necessarily orthogonal) when evaluated at the sample times; functions used for spectral analysis are typically sines and cosines evenly distributed over the frequency range of interest. If we choose too many frequencies in a too-narrow frequency range, the functions will be insufficiently independent, the matrix ill-conditioned, and the resulting spectrum meaningless. [17]

When the basis functions in A are orthogonal (that is, not correlated, meaning the columns have zero pair-wise dot products), the matrix ATA is diagonal; when the columns all have the same power (sum of squares of elements), then that matrix is an identity matrix times a constant, so the inversion is trivial. The latter is the case when the sample times are equally spaced and sinusoids chosen as sines and cosines equally spaced in pairs on the frequency interval 0 to a half cycle per sample (spaced by 1/N cycles per sample, omitting the sine phases at 0 and maximum frequency where they are identically zero). This case is known as the discrete Fourier transform, slightly rewritten in terms of measurements and coefficients. [17]

— DFT case for N equally spaced samples and frequencies, within a scalar factor.

The Lomb method

A power spectrum (magnitude-squared) of two sinusoidal basis functions, calculated by the periodogram method Periodogram.svg
A power spectrum (magnitude-squared) of two sinusoidal basis functions, calculated by the periodogram method

Trying to lower the computational burden of the Vaníček method in 1976 [9] (no longer an issue), Lomb proposed using the above simplification in general, except for pair-wise correlations between sine and cosine bases of the same frequency, since the correlations between pairs of sinusoids are often small, at least when they are not tightly spaced. This formulation is essentially that of the traditional periodogram but adapted for use with unevenly spaced samples. The vector x is a reasonably good estimate of an underlying spectrum, but since we ignore any correlations, Ax is no longer a good approximation to the signal, and the method is no longer a least-squares method — yet in the literature continues to be referred to as such.

Rather than just taking dot products of the data with sine and cosine waveforms directly, Scargle modified the standard periodogram formula so to find a time delay first, such that this pair of sinusoids would be mutually orthogonal at sample times and also adjusted for the potentially unequal powers of these two basis functions, to obtain a better estimate of the power at a frequency. [3] [10] This procedure made his modified periodogram method exactly equivalent to Lomb's method. Time delay by definition equals to

Then the periodogram at frequency is estimated as:

,

which, as Scargle reports, has the same statistical distribution as the periodogram in the evenly sampled case. [10]

At any individual frequency , this method gives the same power as does a least-squares fit to sinusoids of that frequency and of the form:

[18]

In practice, it is always difficult to judge if a given Lomb peak is significant or not, especially when the nature of the noise is unknown, so for example a false-alarm spectral peak in the Lomb periodogram analysis of noisy periodic signal may result from noise in turbulence data. [19] Fourier methods can also report false spectral peaks when analyzing patched-up or data edited otherwise. [7]

The generalized Lomb–Scargle periodogram

The standard Lomb–Scargle periodogram is only valid for a model with a zero mean. Commonly, this is approximated — by subtracting the mean of the data before calculating the periodogram. However, this is an inaccurate assumption when the mean of the model (the fitted sinusoids) is non-zero. The generalized Lomb–Scargle periodogram removes this assumption and explicitly solves for the mean. In this case, the function fitted is

[20]

The generalized Lomb–Scargle periodogram has also been referred to in the literature as a floating mean periodogram. [21]

Korenberg's "fast orthogonal search" method

Michael Korenberg of Queen's University in Kingston, Ontario, developed a method for choosing a sparse set of components from an over-complete set — such as sinusoidal components for spectral analysis — called the fast orthogonal search (FOS). Mathematically, FOS uses a slightly modified Cholesky decomposition in a mean-square error reduction (MSER) process, implemented as a sparse matrix inversion. [15] [22] As with the other LSSA methods, FOS avoids the major shortcoming of discrete Fourier analysis, so it can accurately identify embedded periodicities and excel with unequally spaced data. The fast orthogonal search method was applied to also other problems, such as nonlinear system identification.

Palmer's Chi-squared method

Palmer has developed a method for finding the best-fit function to any chosen number of harmonics, allowing more freedom to find non-sinusoidal harmonic functions. [23] His is a fast (FFT-based) technique for weighted least-squares analysis on arbitrarily spaced data with non-uniform standard errors. Source code that implements this technique is available. [24] Because data are often not sampled at uniformly spaced discrete times, this method "grids" the data by sparsely filling a time series array at the sample times. All intervening grid points receive zero statistical weight, equivalent to having infinite error bars at times between samples.

Applications

Beta distribution for different values of its parameters PDF of the Beta distribution.gif
Beta distribution for different values of its parameters

The most useful feature of LSSA is enabling incomplete records to be spectrally analyzed — without the need to manipulate data or to invent otherwise non-existent data.

Magnitudes in the LSSA spectrum depict the contribution of a frequency or period to the variance of the time series. [4] Generally, spectral magnitudes thus defined enable the output's straightforward significance level regime. [25] Alternatively, spectral magnitudes in the Vaníček spectrum can also be expressed in dB. [26] Note that spectral magnitudes in the Vaníček spectrum follow β-distribution. [27]

Inverse transformation of Vaníček's LSSA is possible, as is most easily seen by writing the forward transform as a matrix; the matrix inverse (when the matrix is not singular) or pseudo-inverse will then be an inverse transformation; the inverse will exactly match the original data if the chosen sinusoids are mutually independent at the sample points and their number is equal to the number of data points. [17] No such inverse procedure is known for the periodogram method.

Implementation

The LSSA can be implemented in less than a page of MATLAB code. [28] In essence: [16]

"to compute the least-squares spectrum we must compute m spectral values ... which involves performing the least-squares approximation m times, each time to get [the spectral power] for a different frequency"

I.e., for each frequency in a desired set of frequencies, sine and cosine functions are evaluated at the times corresponding to the data samples, and dot products of the data vector with the sinusoid vectors are taken and appropriately normalized; following the method known as Lomb/Scargle periodogram, a time shift is calculated for each frequency to orthogonalize the sine and cosine components before the dot product; [17] finally, a power is computed from those two amplitude components. This same process implements a discrete Fourier transform when the data are uniformly spaced in time and the frequencies chosen correspond to integer numbers of cycles over the finite data record.

This method treats each sinusoidal component independently, or out of context, even though they may not be orthogonal to data points; it is Vaníček's original method. In addition, it is possible to perform a full simultaneous or in-context least-squares fit by solving a matrix equation and partitioning the total data variance between the specified sinusoid frequencies. [17] Such a matrix least-squares solution is natively available in MATLAB as the backslash operator. [29]

Furthermore, the simultaneous or in-context method, as opposed to the independent or out-of-context version (as well as the periodogram version due to Lomb), cannot fit more components (sines and cosines) than there are data samples, so that: [17]

"...serious repercussions can also arise if the selected frequencies result in some of the Fourier components (trig functions) becoming nearly linearly dependent with each other, thereby producing an ill-conditioned or near singular N. To avoid such ill conditioning it becomes necessary to either select a different set of frequencies to be estimated (e.g., equally spaced frequencies) or simply neglect the correlations in N (i.e., the off-diagonal blocks) and estimate the inverse least squares transform separately for the individual frequencies..."

Lomb's periodogram method, on the other hand, can use an arbitrarily high number of, or density of, frequency components, as in a standard periodogram; that is, the frequency domain can be over-sampled by an arbitrary factor. [3] However, as mentioned above, one should keep in mind that Lomb's simplification and diverging from the least squares criterion opened up his technique to grave sources of errors, resulting even in false spectral peaks. [19]

In Fourier analysis, such as the Fourier transform and discrete Fourier transform, the sinusoids fitted to data are all mutually orthogonal, so there is no distinction between the simple out-of-context dot-product-based projection onto basis functions versus an in-context simultaneous least-squares fit; that is, no matrix inversion is required to least-squares partition the variance between orthogonal sinusoids of different frequencies. [30] In the past, Fourier's was for many a method of choice thanks to its processing-efficient fast Fourier transform implementation when complete data records with equally spaced samples are available, and they used the Fourier family of techniques to analyze gapped records as well, which, however, required manipulating and even inventing non-existent data just so to be able to run a Fourier-based algorithm.

See also

Related Research Articles

<span class="mw-page-title-main">Discrete Fourier transform</span> Type of Fourier transform in discrete mathematics

In mathematics, the discrete Fourier transform (DFT) converts a finite sequence of equally-spaced samples of a function into a same-length sequence of equally-spaced samples of the discrete-time Fourier transform (DTFT), which is a complex-valued function of frequency. The interval at which the DTFT is sampled is the reciprocal of the duration of the input sequence. An inverse DFT (IDFT) is a Fourier series, using the DTFT samples as coefficients of complex sinusoids at the corresponding DTFT frequencies. It has the same sample-values as the original input sequence. The DFT is therefore said to be a frequency domain representation of the original input sequence. If the original sequence spans all the non-zero values of a function, its DTFT is continuous, and the DFT provides discrete samples of one cycle. If the original sequence is one cycle of a periodic function, the DFT provides all the non-zero values of one DTFT cycle.

<span class="mw-page-title-main">Fourier analysis</span> Branch of mathematics

In mathematics, Fourier analysis is the study of the way general functions may be represented or approximated by sums of simpler trigonometric functions. Fourier analysis grew from the study of Fourier series, and is named after Joseph Fourier, who showed that representing a function as a sum of trigonometric functions greatly simplifies the study of heat transfer.

<span class="mw-page-title-main">Fourier transform</span> Mathematical transform that expresses a function of time as a function of frequency

In physics, engineering and mathematics, the Fourier transform (FT) is an integral transform that converts a function into a form that describes the frequencies present in the original function. The output of the transform is a complex-valued function of frequency. The term Fourier transform refers to both this complex-valued function and the mathematical operation. When a distinction needs to be made the Fourier transform is sometimes called the frequency domain representation of the original function. The Fourier transform is analogous to decomposing the sound of a musical chord into the intensities of its constituent pitches.

In mathematics, the discrete sine transform (DST) is a Fourier-related transform similar to the discrete Fourier transform (DFT), but using a purely real matrix. It is equivalent to the imaginary parts of a DFT of roughly twice the length, operating on real data with odd symmetry (since the Fourier transform of a real and odd function is imaginary and odd), where in some variants the input and/or output data are shifted by half a sample.

<span class="mw-page-title-main">Window function</span> Function used in signal processing

In signal processing and statistics, a window function is a mathematical function that is zero-valued outside of some chosen interval. Typically, windows functions are symmetric around the middle of the interval, approach a maximum in the middle, and taper away from the middle. Mathematically, when another function or waveform/data-sequence is "multiplied" by a window function, the product is also zero-valued outside the interval: all that is left is the part where they overlap, the "view through the window". Equivalently, and in actual practice, the segment of data within the window is first isolated, and then only that data is multiplied by the window function values. Thus, tapering, not segmentation, is the main purpose of window functions.

<span class="mw-page-title-main">Sine wave</span> Wave shaped like the sine function

A sine wave, sinusoidal wave, or sinusoid is a periodic wave whose waveform (shape) is the trigonometric sine function. In mechanics, as a linear motion over time, this is simple harmonic motion; as rotation, it corresponds to uniform circular motion. Sine waves occur often in physics, including wind waves, sound waves, and light waves, such as monochromatic radiation. In engineering, signal processing, and mathematics, Fourier analysis decomposes general functions into a sum of sine waves of various frequencies, relative phases, and magnitudes.

<span class="mw-page-title-main">Short-time Fourier transform</span> Fourier-related transform suited to signals that change rather quickly in time

The short-time Fourier transform (STFT), is a Fourier-related transform used to determine the sinusoidal frequency and phase content of local sections of a signal as it changes over time. In practice, the procedure for computing STFTs is to divide a longer time signal into shorter segments of equal length and then compute the Fourier transform separately on each shorter segment. This reveals the Fourier spectrum on each shorter segment. One then usually plots the changing spectra as a function of time, known as a spectrogram or waterfall plot, such as commonly used in software defined radio (SDR) based spectrum displays. Full bandwidth displays covering the whole range of an SDR commonly use fast Fourier transforms (FFTs) with 2^24 points on desktop computers.

In signal processing, a periodogram is an estimate of the spectral density of a signal. The term was coined by Arthur Schuster in 1898. Today, the periodogram is a component of more sophisticated methods. It is the most common tool for examining the amplitude vs frequency characteristics of FIR filters and window functions. FFT spectrum analyzers are also implemented as a time-sequence of periodograms.

The Fourier transform of a function of time, s(t), is a complex-valued function of frequency, S(f), often referred to as a frequency spectrum. Any linear time-invariant operation on s(t) produces a new spectrum of the form H(f)•S(f), which changes the relative magnitudes and/or angles (phase) of the non-zero values of S(f). Any other type of operation creates new frequency components that may be referred to as spectral leakage in the broadest sense. Sampling, for instance, produces leakage, which we call aliases of the original spectral component. For Fourier transform purposes, sampling is modeled as a product between s(t) and a Dirac comb function. The spectrum of a product is the convolution between S(f) and another function, which inevitably creates the new frequency components. But the term 'leakage' usually refers to the effect of windowing, which is the product of s(t) with a different kind of function, the window function. Window functions happen to have finite duration, but that is not necessary to create leakage. Multiplication by a time-variant function is sufficient.

In mathematics, the Hartley transform (HT) is an integral transform closely related to the Fourier transform (FT), but which transforms real-valued functions to real-valued functions. It was proposed as an alternative to the Fourier transform by Ralph V. L. Hartley in 1942, and is one of many known Fourier-related transforms. Compared to the Fourier transform, the Hartley transform has the advantages of transforming real functions to real functions and of being its own inverse.

<span class="mw-page-title-main">Negative frequency</span>

In mathematics, signed frequency expands upon the concept of frequency, from just an absolute value representing how often some repeating event occurs, to also have a positive or negative sign representing one of two opposing orientations for occurrences of those events. The following examples help illustrate the concept:

In mathematics, the discrete-time Fourier transform (DTFT) is a form of Fourier analysis that is applicable to a sequence of discrete values.

<span class="mw-page-title-main">Prony's method</span> Method to estimate the components of a signal

Prony analysis was developed by Gaspard Riche de Prony in 1795. However, practical use of the method awaited the digital computer. Similar to the Fourier transform, Prony's method extracts valuable information from a uniformly sampled signal and builds a series of damped complex exponentials or damped sinusoids. This allows the estimation of frequency, amplitude, phase and damping components of a signal.

<span class="mw-page-title-main">MUSIC (algorithm)</span> Algorithm used for frequency estimation and radio direction finding

MUSIC is an algorithm used for frequency estimation and radio direction finding.

Geophysical survey is the systematic collection of geophysical data for spatial studies. Detection and analysis of the geophysical signals forms the core of Geophysical signal processing. The magnetic and gravitational fields emanating from the Earth's interior hold essential information concerning seismic activities and the internal structure. Hence, detection and analysis of the electric and Magnetic fields is very crucial. As the Electromagnetic and gravitational waves are multi-dimensional signals, all the 1-D transformation techniques can be extended for the analysis of these signals as well. Hence this article also discusses multi-dimensional signal processing techniques.

In statistical signal processing, the goal of spectral density estimation (SDE) or simply spectral estimation is to estimate the spectral density of a signal from a sequence of time samples of the signal. Intuitively speaking, the spectral density characterizes the frequency content of the signal. One purpose of estimating the spectral density is to detect any periodicities in the data, by observing peaks at the frequencies corresponding to these periodicities.

In statistics, signal processing, and time series analysis, a sinusoidal model is used to approximate a sequence Yi to a sine function:

SigSpec is a statistical technique to provide the reliability of periodicities in a measured time series. It relies on the amplitude spectrum obtained by the Discrete Fourier transform (DFT) and assigns a quantity called the spectral significance to each amplitude. This quantity is a logarithmic measure of the probability that the given amplitude level would be seen in white noise, in the sense of a type I error. It represents the answer to the question, “What would be the chance to obtain an amplitude like the measured one or higher, if the analysed time series were random?”

<span class="mw-page-title-main">Frequency selective surface</span> Optical filter

A frequency-selective surface (FSS) is any thin, repetitive surface designed to reflect, transmit or absorb electromagnetic fields based on the frequency of the field. In this sense, an FSS is a type of optical filter or metal-mesh optical filters in which the filtering is accomplished by virtue of the regular, periodic pattern on the surface of the FSS. Though not explicitly mentioned in the name, FSS's also have properties which vary with incidence angle and polarization as well - these are unavoidable consequences of the way in which FSS's are constructed. Frequency-selective surfaces have been most commonly used in the radio frequency region of the electromagnetic spectrum and find use in applications as diverse as the aforementioned microwave oven, antenna radomes and modern metamaterials. Sometimes frequency selective surfaces are referred to simply as periodic surfaces and are a 2-dimensional analog of the new periodic volumes known as photonic crystals.

Multiresolution Fourier Transform is an integral fourier transform that represents a specific wavelet-like transform with a fully scalable modulated window, but not all possible translations.

References

  1. Cafer Ibanoglu (2000). Variable Stars As Essential Astrophysical Tools. Springer. ISBN   0-7923-6084-2.
  2. 1 2 D. Scott Birney; David Oesper; Guillermo Gonzalez (2006). Observational Astronomy. Cambridge University Press. ISBN   0-521-85370-2.
  3. 1 2 3 4 5 Press (2007). Numerical Recipes (3rd ed.). Cambridge University Press. ISBN   978-0-521-88068-8.
  4. 1 2 3 P. Vaníček (1 August 1969). "Approximate Spectral Analysis by Least-squares Fit" (PDF). Astrophysics and Space Science. 4 (4): 387–391. Bibcode:1969Ap&SS...4..387V. doi:10.1007/BF00651344. OCLC   5654872875. S2CID   124921449.
  5. 1 2 P. Vaníček (1 July 1971). "Further development and properties of the spectral analysis by least-squares fit" (PDF). Astrophysics and Space Science. 12 (1): 10–33. Bibcode:1971Ap&SS..12...10V. doi:10.1007/BF00656134. S2CID   109404359.
  6. J. Taylor; S. Hamilton (20 March 1972). "Some tests of the Vaníček Method of spectral analysis". Astrophysics and Space Science. 17 (2): 357–367. Bibcode:1972Ap&SS..17..357T. doi:10.1007/BF00642907. S2CID   123569059.
  7. 1 2 M. Omerbashich (26 June 2006). "Gauss-Vanicek spectral analysis of the Sepkoski compendium: no new life cycles". Computing in Science & Engineering. 8 (4): 26–30. arXiv: math-ph/0608014 . Bibcode:2006CSE.....8d..26O. doi:10.1109/MCSE.2006.68.
  8. Hans P. A. Van Dongen (1999). "Searching for Biological Rhythms: Peak Detection in the Periodogram of Unequally Spaced Data". Journal of Biological Rhythms. 14 (6): 617–620. doi:10.1177/074873099129000984. PMID   10643760. S2CID   14886901.
  9. 1 2 3 Lomb, N. R. (1976). "Least-squares frequency analysis of unequally spaced data". Astrophysics and Space Science. 39 (2): 447–462. Bibcode:1976Ap&SS..39..447L. doi:10.1007/BF00648343. S2CID   2671466.
  10. 1 2 3 4 5 Scargle, J. D. (1982). "Studies in astronomical time series analysis. II - Statistical aspects of spectral analysis of unevenly spaced data". Astrophysical Journal. 263: 835. Bibcode:1982ApJ...263..835S. doi:10.1086/160554.
  11. David Brunt (1931). The Combination of Observations (2nd ed.). Cambridge University Press.
  12. Barning, F. J. M. (1963). "The numerical analysis of the light-curve of 12 Lacertae". Bulletin of the Astronomical Institutes of the Netherlands. 17: 22. Bibcode:1963BAN....17...22B.
  13. 1 2 Pascal Vincent; Yoshua Bengio (2002). "Kernel Matching Pursuit" (PDF). Machine Learning. 48: 165–187. doi: 10.1023/A:1013955821559 .
  14. Y. C. Pati, R. Rezaiifar, and P. S. Krishnaprasad, "Orthogonal matching pursuit: Recursive function approximation with applications to wavelet decomposition," in Proc. 27th Asilomar Conference on Signals, Systems and Computers, A. Singh, ed., Los Alamitos, CA, USA, IEEE Computer Society Press, 1993
  15. 1 2 Korenberg, M. J. (1989). "A robust orthogonal algorithm for system identification and time-series analysis". Biological Cybernetics. 60 (4): 267–276. doi:10.1007/BF00204124. PMID   2706281. S2CID   11712196.
  16. 1 2 Wells, D.E., P. Vaníček, S. Pagiatakis, 1985. Least-squares spectral analysis revisited. Department of Surveying Engineering Technical Report 84, University of New Brunswick, Fredericton, 68 pages, Available at .
  17. 1 2 3 4 5 6 7 Craymer, M.R., The Least Squares Spectrum, Its Inverse Transform and Autocorrelation Function: Theory and Some Applications in Geodesy, Ph.D. Dissertation, University of Toronto, Canada (1998).
  18. William J. Emery; Richard E. Thomson (2001). Data Analysis Methods in Physical Oceanography. Elsevier. ISBN   0-444-50756-6.
  19. 1 2 Zhou, W.-X.; Sornette, D. (October 2001). "Statistical significance of periodicity and log-periodicity with heavy-tailed correlated noise". International Journal of Modern Physics C. 13 (2): 137–169. arXiv: cond-mat/0110445 . Bibcode:2002IJMPC..13..137Z. doi:10.1142/S0129183102003024. S2CID   8256563.
  20. M. Zechmeister; M. Kürster (March 2009). "The generalised Lomb–Scargle periodogram. A new formalism for the floating-mean and Keplerian periodograms". Astronomy & Astrophysics. 496 (2): 577–584. arXiv: 0901.2573 . Bibcode:2009A&A...496..577Z. doi:10.1051/0004-6361:200811296. S2CID   10408194.
  21. Andrew Cumming; Geoffrey W. Marcy; R. Paul Butler (December 1999). "The Lick Planet Search: Detectability and Mass Thresholds". The Astrophysical Journal. 526 (2): 890–915. arXiv: astro-ph/9906466 . Bibcode:1999ApJ...526..890C. doi:10.1086/308020. S2CID   12560512.
  22. Korenberg, Michael J.; Brenan, Colin J. H.; Hunter, Ian W. (1997). "Raman Spectral Estimation via Fast Orthogonal Search". The Analyst. 122 (9): 879–882. Bibcode:1997Ana...122..879K. doi:10.1039/a700902j.
  23. Palmer, David M. (2009). "A Fast Chi-squared Technique For Period Search of Irregularly Sampled Data". The Astrophysical Journal. 695 (1): 496–502. arXiv: 0901.1913 . Bibcode:2009ApJ...695..496P. doi:10.1088/0004-637X/695/1/496. S2CID   5991300.
  24. "David Palmer: The Fast Chi-squared Period Search".
  25. Beard, A.G., Williams, P.J.S., Mitchell, N.J. & Muller, H.G. A special climatology of planetary waves and tidal variability, J Atm. Solar-Ter. Phys. 63 (09), p.801–811 (2001).
  26. Pagiatakis, S. Stochastic significance of peaks in the least-squares spectrum, J of Geodesy 73, p.67-78 (1999).
  27. Steeves, R.R. A statistical test for significance of peaks in the least squares spectrum, Collected Papers of the Geodetic Survey, Department of Energy, Mines and Resources, Surveys and Mapping, Ottawa, Canada, p.149-166 (1981)
  28. Richard A. Muller; Gordon J. MacDonald (2000). Ice Ages and Astronomical Causes: Data, spectral analysis and mechanisms (1st ed.). Springer Berlin Heidelberg. Bibcode:2000iaac.book.....M. ISBN   978-3-540-43779-6. OL   20645181M. Wikidata   Q111312009.
  29. Timothy A. Davis; Kermit Sigmon (2005). MATLAB Primer. CRC Press. ISBN   1-58488-523-8.
  30. Darrell Williamson (1999). Discrete-Time Signal Processing: An Algebraic Approach. Springer. ISBN   1-85233-161-5.