Matched filter

Last updated

In signal processing, the output of the matched filter is given by correlating a known delayed signal, or template, with an unknown signal to detect the presence of the template in the unknown signal. [1] [2] This is equivalent to convolving the unknown signal with a conjugated time-reversed version of the template. The matched filter is the optimal linear filter for maximizing the signal-to-noise ratio (SNR) in the presence of additive stochastic noise.

Contents

Matched filters are commonly used in radar, in which a known signal is sent out, and the reflected signal is examined for common elements of the out-going signal. Pulse compression is an example of matched filtering. It is so called because the impulse response is matched to input pulse signals. Two-dimensional matched filters are commonly used in image processing, e.g., to improve the SNR of X-ray observations.

Matched filtering is a demodulation technique with LTI (linear time invariant) filters to maximize SNR. [3] It was originally also known as a North filter. [4]

Derivation

Derivation via matrix algebra

The following section derives the matched filter for a discrete-time system. The derivation for a continuous-time system is similar, with summations replaced with integrals.

The matched filter is the linear filter, , that maximizes the output signal-to-noise ratio.

where is the input as a function of the independent variable , and is the filtered output. Though we most often express filters as the impulse response of convolution systems, as above (see LTI system theory), it is easiest to think of the matched filter in the context of the inner product, which we will see shortly.

We can derive the linear filter that maximizes output signal-to-noise ratio by invoking a geometric argument. The intuition behind the matched filter relies on correlating the received signal (a vector) with a filter (another vector) that is parallel with the signal, maximizing the inner product. This enhances the signal. When we consider the additive stochastic noise, we have the additional challenge of minimizing the output due to noise by choosing a filter that is orthogonal to the noise.

Let us formally define the problem. We seek a filter, , such that we maximize the output signal-to-noise ratio, where the output is the inner product of the filter and the observed signal .

Our observed signal consists of the desirable signal and additive noise :

Let us define the auto-correlation matrix of the noise, reminding ourselves that this matrix has Hermitian symmetry, a property that will become useful in the derivation:

where denotes the conjugate transpose of , and denotes expectation (note that in case the noise has zero-mean, its auto-correlation matrix is equal to its covariance matrix).

Let us call our output, , the inner product of our filter and the observed signal such that

We now define the signal-to-noise ratio, which is our objective function, to be the ratio of the power of the output due to the desired signal to the power of the output due to the noise:

We rewrite the above:

We wish to maximize this quantity by choosing . Expanding the denominator of our objective function, we have

Now, our becomes

We will rewrite this expression with some matrix manipulation. The reason for this seemingly counterproductive measure will become evident shortly. Exploiting the Hermitian symmetry of the auto-correlation matrix , we can write

We would like to find an upper bound on this expression. To do so, we first recognize a form of the Cauchy–Schwarz inequality:

which is to say that the square of the inner product of two vectors can only be as large as the product of the individual inner products of the vectors. This concept returns to the intuition behind the matched filter: this upper bound is achieved when the two vectors and are parallel. We resume our derivation by expressing the upper bound on our in light of the geometric inequality above:

Our valiant matrix manipulation has now paid off. We see that the expression for our upper bound can be greatly simplified:

We can achieve this upper bound if we choose,

where is an arbitrary real number. To verify this, we plug into our expression for the output :

Thus, our optimal matched filter is

We often choose to normalize the expected value of the power of the filter output due to the noise to unity. That is, we constrain

This constraint implies a value of , for which we can solve:

yielding

giving us our normalized filter,

If we care to write the impulse response of the filter for the convolution system, it is simply the complex conjugate time reversal of the input .

Though we have derived the matched filter in discrete time, we can extend the concept to continuous-time systems if we replace with the continuous-time autocorrelation function of the noise, assuming a continuous signal , continuous noise , and a continuous filter .

Derivation via Lagrangian

Alternatively, we may solve for the matched filter by solving our maximization problem with a Lagrangian. Again, the matched filter endeavors to maximize the output signal-to-noise ratio () of a filtered deterministic signal in stochastic additive noise. The observed sequence, again, is

with the noise auto-correlation matrix,

The signal-to-noise ratio is

where and .

Evaluating the expression in the numerator, we have

and in the denominator,

The signal-to-noise ratio becomes

If we now constrain the denominator to be 1, the problem of maximizing is reduced to maximizing the numerator. We can then formulate the problem using a Lagrange multiplier:

which we recognize as a generalized eigenvalue problem

Since is of unit rank, it has only one nonzero eigenvalue. It can be shown that this eigenvalue equals

yielding the following optimal matched filter

This is the same result found in the previous subsection.

Interpretation as a least-squares estimator

Derivation

Matched filtering can also be interpreted as a least-squares estimator for the optimal location and scaling of a given model or template. Once again, let the observed sequence be defined as

where is uncorrelated zero mean noise. The signal is assumed to be a scaled and shifted version of a known model sequence :

We want to find optimal estimates and for the unknown shift and scaling by minimizing the least-squares residual between the observed sequence and a "probing sequence" :

The appropriate will later turn out to be the matched filter, but is as yet unspecified. Expanding and the square within the sum yields

The first term in brackets is a constant (since the observed signal is given) and has no influence on the optimal solution. The last term has constant expected value because the noise is uncorrelated and has zero mean. We can therefore drop both terms from the optimization. After reversing the sign, we obtain the equivalent optimization problem

Setting the derivative w.r.t. to zero gives an analytic solution for :

Inserting this into our objective function yields a reduced maximization problem for just :

The numerator can be upper-bounded by means of the Cauchy–Schwarz inequality:

The optimization problem assumes its maximum when equality holds in this expression. According to the properties of the Cauchy–Schwarz inequality, this is only possible when

for arbitrary non-zero constants or , and the optimal solution is obtained at as desired. Thus, our "probing sequence" must be proportional to the signal model , and the convenient choice yields the matched filter

Note that the filter is the mirrored signal model. This ensures that the operation to be applied in order to find the optimum is indeed the convolution between the observed sequence and the matched filter . The filtered sequence assumes its maximum at the position where the observed sequence best matches (in a least-squares sense) the signal model .

Implications

The matched filter may be derived in a variety of ways, [2] but as a special case of a least-squares procedure it may also be interpreted as a maximum likelihood method in the context of a (coloured) Gaussian noise model and the associated Whittle likelihood. [5] If the transmitted signal possessed no unknown parameters (like time-of-arrival, amplitude,...), then the matched filter would, according to the Neyman–Pearson lemma, minimize the error probability. However, since the exact signal generally is determined by unknown parameters that effectively are estimated (or fitted) in the filtering process, the matched filter constitutes a generalized maximum likelihood (test-) statistic. [6] The filtered time series may then be interpreted as (proportional to) the profile likelihood, the maximized conditional likelihood as a function of the time parameter. [7] This implies in particular that the error probability (in the sense of Neyman and Pearson, i.e., concerning maximization of the detection probability for a given false-alarm probability [8] ) is not necessarily optimal. What is commonly referred to as the Signal-to-noise ratio (SNR) , which is supposed to be maximized by a matched filter, in this context corresponds to , where is the (conditionally) maximized likelihood ratio. [7] [nb 1]

The construction of the matched filter is based on a known noise spectrum. In reality, however, the noise spectrum is usually estimated from data and hence only known up to a limited precision. For the case of an uncertain spectrum, the matched filter may be generalized to a more robust iterative procedure with favourable properties also in non-Gaussian noise. [7]

Frequency-domain interpretation

When viewed in the frequency domain, it is evident that the matched filter applies the greatest weighting to spectral components exhibiting the greatest signal-to-noise ratio (i.e., large weight where noise is relatively low, and vice versa). In general this requires a non-flat frequency response, but the associated "distortion" is no cause for concern in situations such as radar and digital communications, where the original waveform is known and the objective is the detection of this signal against the background noise. On the technical side, the matched filter is a weighted least-squares method based on the (heteroscedastic) frequency-domain data (where the "weights" are determined via the noise spectrum, see also previous section), or equivalently, a least-squares method applied to the whitened data.

Examples

Radar and sonar

Matched filters are often used in signal detection. [1] As an example, suppose that we wish to judge the distance of an object by reflecting a signal off it. We may choose to transmit a pure-tone sinusoid at 1 Hz. We assume that our received signal is an attenuated and phase-shifted form of the transmitted signal with added noise.

To judge the distance of the object, we correlate the received signal with a matched filter, which, in the case of white (uncorrelated) noise, is another pure-tone 1-Hz sinusoid. When the output of the matched filter system exceeds a certain threshold, we conclude with high probability that the received signal has been reflected off the object. Using the speed of propagation and the time that we first observe the reflected signal, we can estimate the distance of the object. If we change the shape of the pulse in a specially-designed way, the signal-to-noise ratio and the distance resolution can be even improved after matched filtering: this is a technique known as pulse compression.

Additionally, matched filters can be used in parameter estimation problems (see estimation theory). To return to our previous example, we may desire to estimate the speed of the object, in addition to its position. To exploit the Doppler effect, we would like to estimate the frequency of the received signal. To do so, we may correlate the received signal with several matched filters of sinusoids at varying frequencies. The matched filter with the highest output will reveal, with high probability, the frequency of the reflected signal and help us determine the radial velocity of the object, i.e. the relative speed either directly towards or away from the observer. This method is, in fact, a simple version of the discrete Fourier transform (DFT). The DFT takes an -valued complex input and correlates it with matched filters, corresponding to complex exponentials at different frequencies, to yield complex-valued numbers corresponding to the relative amplitudes and phases of the sinusoidal components (see Moving target indication).

Digital communications

The matched filter is also used in communications. In the context of a communication system that sends binary messages from the transmitter to the receiver across a noisy channel, a matched filter can be used to detect the transmitted pulses in the noisy received signal.

Matched Filter Total System.jpg

Imagine we want to send the sequence "0101100100" coded in non polar non-return-to-zero (NRZ) through a certain channel.

Mathematically, a sequence in NRZ code can be described as a sequence of unit pulses or shifted rect functions, each pulse being weighted by +1 if the bit is "1" and by -1 if the bit is "0". Formally, the scaling factor for the bit is,

We can represent our message, , as the sum of shifted unit pulses:

where is the time length of one bit and is the rectangular function.

Thus, the signal to be sent by the transmitter is

Original message.svg

If we model our noisy channel as an AWGN channel, white Gaussian noise is added to the signal. At the receiver end, for a Signal-to-noise ratio of 3 dB, this may look like:

Received message.svg

A first glance will not reveal the original transmitted sequence. There is a high power of noise relative to the power of the desired signal (i.e., there is a low signal-to-noise ratio). If the receiver were to sample this signal at the correct moments, the resulting binary message could be incorrect.

To increase our signal-to-noise ratio, we pass the received signal through a matched filter. In this case, the filter should be matched to an NRZ pulse (equivalent to a "1" coded in NRZ code). Precisely, the impulse response of the ideal matched filter, assuming white (uncorrelated) noise should be a time-reversed complex-conjugated scaled version of the signal that we are seeking. We choose

In this case, due to symmetry, the time-reversed complex conjugate of is in fact , allowing us to call the impulse response of our matched filter convolution system.

After convolving with the correct matched filter, the resulting signal, is,

where denotes convolution.

Filtered message.svg

Which can now be safely sampled by the receiver at the correct sampling instants, and compared to an appropriate threshold, resulting in a correct interpretation of the binary message.

Filtered message threshold.svg

Gravitational-wave astronomy

Matched filters play a central role in gravitational-wave astronomy. [9] The first observation of gravitational waves was based on large-scale filtering of each detector's output for signals resembling the expected shape, followed by subsequent screening for coincident and coherent triggers between both instruments. [10] False-alarm rates, and with that, the statistical significance of the detection were then assessed using resampling methods. [11] [12] Inference on the astrophysical source parameters was completed using Bayesian methods based on parameterized theoretical models for the signal waveform and (again) on the Whittle likelihood. [13] [14]

Biology

Animals living in relatively static environments would have relatively fixed features of the environment to perceive. This allows the evolution of filters that match the expected signal with the highest signal-to-noise ratio, the matched filter. [15] Sensors that perceive the world "through such a 'matched filter' severely limits the amount of information the brain can pick up from the outside world, but it frees the brain from the need to perform more intricate computations to extract the information finally needed for fulfilling a particular task." [16]

See also

Notes

  1. The common reference to SNR has in fact been criticized as somewhat misleading: "The interesting feature of this approach is that theoretical perfection is attained without aiming consciously at a maximum signal/noise ratio. As the matter of quite incidental interest, it happens that the operation [...] does maximize the peak signal/noise ratio, but this fact plays no part whatsoever in the present theory. Signal/noise ratio is not a measure of information [...]." (Woodward, 1953; [1] Sec.5.1).

Related Research Articles

Signal-to-noise ratio is a measure used in science and engineering that compares the level of a desired signal to the level of background noise. SNR is defined as the ratio of signal power to noise power, often expressed in decibels. A ratio higher than 1:1 indicates more signal than noise.

The total harmonic distortion is a measurement of the harmonic distortion present in a signal and is defined as the ratio of the sum of the powers of all harmonic components to the power of the fundamental frequency. Distortion factor, a closely related term, is sometimes used as a synonym.

In statistics, maximum likelihood estimation (MLE) is a method of estimating the parameters of an assumed probability distribution, given some observed data. This is achieved by maximizing a likelihood function so that, under the assumed statistical model, the observed data is most probable. The point in the parameter space that maximizes the likelihood function is called the maximum likelihood estimate. The logic of maximum likelihood is both intuitive and flexible, and as such the method has become a dominant means of statistical inference.

In mathematical analysis, Hölder's inequality, named after Otto Hölder, is a fundamental inequality between integrals and an indispensable tool for the study of Lp spaces.

In mathematical analysis, the Minkowski inequality establishes that the Lp spaces are normed vector spaces. Let be a measure space, let and let and be elements of Then is in and we have the triangle inequality

Additive white Gaussian noise (AWGN) is a basic noise model used in information theory to mimic the effect of many random processes that occur in nature. The modifiers denote specific characteristics:

<span class="mw-page-title-main">Thermodynamic potential</span> Scalar physical quantities representing system states

A thermodynamic potential is a scalar quantity used to represent the thermodynamic state of a system. Just as in mechanics, where potential energy is defined as capacity to do work, similarly different potentials have different meanings. The concept of thermodynamic potentials was introduced by Pierre Duhem in 1886. Josiah Willard Gibbs in his papers used the term fundamental functions.

<span class="mw-page-title-main">Helmholtz free energy</span> Thermodynamic potential

In thermodynamics, the Helmholtz free energy is a thermodynamic potential that measures the useful work obtainable from a closed thermodynamic system at a constant temperature (isothermal). The change in the Helmholtz energy during a process is equal to the maximum amount of work that the system can perform in a thermodynamic process in which temperature is held constant. At constant temperature, the Helmholtz free energy is minimized at equilibrium.

In the calculus of variations and classical mechanics, the Euler–Lagrange equations are a system of second-order ordinary differential equations whose solutions are stationary points of the given action functional. The equations were discovered in the 1750s by Swiss mathematician Leonhard Euler and Italian mathematician Joseph-Louis Lagrange.

Similarity may refer to:

Linear discriminant analysis (LDA), normal discriminant analysis (NDA), or discriminant function analysis is a generalization of Fisher's linear discriminant, a method used in statistics and other fields, to find a linear combination of features that characterizes or separates two or more classes of objects or events. The resulting combination may be used as a linear classifier, or, more commonly, for dimensionality reduction before later classification.

In statistics and information theory, a maximum entropy probability distribution has entropy that is at least as great as that of all other members of a specified class of probability distributions. According to the principle of maximum entropy, if nothing is known about a distribution except that it belongs to a certain class, then the distribution with the largest entropy should be chosen as the least-informative default. The motivation is twofold: first, maximizing entropy minimizes the amount of prior information built into the distribution; second, many physical systems tend to move towards maximal entropy configurations over time.

<span class="mw-page-title-main">Rice distribution</span> Probability distribution

In probability theory, the Rice distribution or Rician distribution is the probability distribution of the magnitude of a circularly-symmetric bivariate normal random variable, possibly with non-zero mean (noncentral). It was named after Stephen O. Rice (1907–1986).

Least mean squares (LMS) algorithms are a class of adaptive filter used to mimic a desired filter by finding the filter coefficients that relate to producing the least mean square of the error signal. It is a stochastic gradient descent method in that the filter is only adapted based on the error at the current time. It was invented in 1960 by Stanford University professor Bernard Widrow and his first Ph.D. student, Ted Hoff, based on their research in single-layer neural networks (ADALINE). Specifically, they used gradient descent to train ADALINE to recognize patterns, and called the algorithm "delta rule". They then applied the rule to filters, resulting in the LMS algorithm.

<span class="mw-page-title-main">Wiener deconvolution</span>

In mathematics, Wiener deconvolution is an application of the Wiener filter to the noise problems inherent in deconvolution. It works in the frequency domain, attempting to minimize the impact of deconvolved noise at frequencies which have a poor signal-to-noise ratio.

Pulse compression is a signal processing technique commonly used by radar, sonar and echography to either increase the range resolution when pulse length is constrained or increase the signal to noise ratio when the peak power and the bandwidth of the transmitted signal are constrained. This is achieved by modulating the transmitted pulse and then correlating the received signal with the transmitted pulse.

Precoding is a generalization of beamforming to support multi-stream transmission in multi-antenna wireless communications. In conventional single-stream beamforming, the same signal is emitted from each of the transmit antennas with appropriate weighting such that the signal power is maximized at the receiver output. When the receiver has multiple antennas, single-stream beamforming cannot simultaneously maximize the signal level at all of the receive antennas. In order to maximize the throughput in multiple receive antenna systems, multi-stream transmission is generally required.

In nonideal fluid dynamics, the Hagen–Poiseuille equation, also known as the Hagen–Poiseuille law, Poiseuille law or Poiseuille equation, is a physical law that gives the pressure drop in an incompressible and Newtonian fluid in laminar flow flowing through a long cylindrical pipe of constant cross section. It can be successfully applied to air flow in lung alveoli, or the flow through a drinking straw or through a hypodermic needle. It was experimentally derived independently by Jean Léonard Marie Poiseuille in 1838 and Gotthilf Heinrich Ludwig Hagen, and published by Poiseuille in 1840–41 and 1846. The theoretical justification of the Poiseuille law was given by George Stokes in 1845.

Signal averaging is a signal processing technique applied in the time domain, intended to increase the strength of a signal relative to noise that is obscuring it. By averaging a set of replicate measurements, the signal-to-noise ratio (SNR) will be increased, ideally in proportion to the square root of the number of measurements.

In probability theory, the family of complex normal distributions, denoted or , characterizes complex random variables whose real and imaginary parts are jointly normal. The complex normal family has three parameters: location parameter μ, covariance matrix , and the relation matrix . The standard complex normal is the univariate distribution with , , and .

References

  1. 1 2 3 Woodward, P. M. (1953). Probability and information theory with applications to radar. London: Pergamon Press.
  2. 1 2 Turin, G. L. (1960). "An introduction to matched filters". IRE Transactions on Information Theory. 6 (3): 311–329. doi:10.1109/TIT.1960.1057571. S2CID   5128742.
  3. "Demodulation". OpenStax CNX. Retrieved 2017-04-18.
  4. After D.O. North who was among the first to introduce the concept: North, D. O. (1943). "An analysis of the factors which determine signal/noise discrimination in pulsed carrier systems". Report PPR-6C, RCA Laboratories, Princeton, NJ.
    Re-print: North, D. O. (1963). "An analysis of the factors which determine signal/noise discrimination in pulsed-carrier systems". Proceedings of the IEEE. 51 (7): 1016–1027. doi:10.1109/PROC.1963.2383.
    See also: Jaynes, E. T. (2003). "14.6.1 The classical matched filter". Probability theory: The logic of science. Cambridge: Cambridge University Press.
  5. Choudhuri, N.; Ghosal, S.; Roy, A. (2004). "Contiguity of the Whittle measure for a Gaussian time series". Biometrika. 91 (4): 211–218. doi: 10.1093/biomet/91.1.211 .
  6. Mood, A. M.; Graybill, F. A.; Boes, D. C. (1974). "IX. Tests of hypotheses". Introduction to the theory of statistics (3rd ed.). New York: McGraw-Hill.
  7. 1 2 3 Röver, C. (2011). "Student-t based filter for robust signal detection". Physical Review D. 84 (12): 122004. arXiv: 1109.0442 . Bibcode:2011PhRvD..84l2004R. doi:10.1103/PhysRevD.84.122004.
  8. Neyman, J.; Pearson, E. S. (1933). "On the problem of the most efficient tests of statistical hypotheses". Philosophical Transactions of the Royal Society of London A. 231 (694–706): 289–337. Bibcode:1933RSPTA.231..289N. doi: 10.1098/rsta.1933.0009 .
  9. Schutz, B. F. (1999). "Gravitational wave astronomy". Classical and Quantum Gravity. 16 (12A): A131–A156. arXiv: gr-qc/9911034 . Bibcode:1999CQGra..16A.131S. doi:10.1088/0264-9381/16/12A/307. S2CID   19021009.
  10. "LIGO: How We Searched For Merging Black Holes And Found GW150914". A technique known as matched filtering is used to see if there are any signals contained within our data. The aim of matched filtering is to see if the data contains any signals similar to a template bank member. Since our templates should describe the gravitational waveforms for the range of different merging systems that we expect to be able to see, any sufficiently loud signal should be found by this method.
  11. Usman, Samantha A. (2016). "The PyCBC search for gravitational waves from compact binary coalescence". Class. Quantum Grav. 33 (21): 215004. arXiv: 1508.02357 . Bibcode:2016CQGra..33u5004U. doi:10.1088/0264-9381/33/21/215004. S2CID   53979477.
  12. Abbott, B. P.; et al. (The LIGO Scientific Collaboration, the Virgo Collaboration) (2016). "GW150914: First results from the search for binary black hole coalescence with Advanced LIGO". Physical Review D. 93 (12): 122003. arXiv: 1602.03839 . Bibcode:2016PhRvD..93l2003A. doi:10.1103/PhysRevD.93.122003. PMC   7430253 . PMID   32818163.
  13. Abbott, B. P.; et al. (The LIGO Scientific Collaboration, the Virgo Collaboration) (2016). "Properties of the binary black hole merger GW150914". Physical Review Letters. 116 (24): 241102. arXiv: 1602.03840 . Bibcode:2016PhRvL.116x1102A. doi:10.1103/PhysRevLett.116.241102. PMID   27367378. S2CID   217406416.
  14. Meyer, R.; Christensen, N. (2016). "Gravitational waves: A statistical autopsy of a black hole merger". Significance. 13 (2): 20–25. doi: 10.1111/j.1740-9713.2016.00896.x .
  15. Warrant, Eric J. (October 2016). "Sensory matched filters". Current Biology. 26 (20): R976–R980. doi: 10.1016/j.cub.2016.05.042 . ISSN   0960-9822. PMID   27780072.
  16. Wehner, Rüdiger (1987). "'Matched filters': neural models of the external world". Journal of Comparative Physiology A. 161 (4): 511–531. doi:10.1007/bf00603659. ISSN   0340-7594. S2CID   32779686.

Further reading