11
Applications of FIR State Estimators

Design is not how it looks like and feels like. Design is how it works.

Steve Jobs (Apple Inc. cofounder and former CEO)

This chapter is the last, and it is now a proper place to give examples of practical applications of FIR state estimators. As the approach suggests, many useful FIR engineering algorithms can be developed to solve filtering, smoothing, and prediction problems on data finite horizons under diverse operation conditions in different environments. For Gaussian processes, the batch OFIR state estimator has proven to be the most accurate in the MSE sense, and the supporting recursive KF algorithm and its various modifications have found a huge number of applications. The OUFIR state estimator and RH MVF are commonly used in the canonical ML FIR batch form, since the recursive forms for these batches are more complex than Kalman recursions. The UFIR state estimator is blind on optimal horizons (has no other tuning factors) and is thus robust, unlike the OFIR and OUFIR estimators. Therefore, the iterative UFIR filtering algorithm using recursions has found practical applications as an alternative to KF in uncertain environments. It is worth noting that there has been no further development of the original LMF idea, because LMF is nothing more than KF operating on data finite horizons. In recent decades, there has appeared a big class of norm‐bounded upper H 2, upper H Subscript infinity, script upper L 2‐to‐script upper L Subscript infinity, script upper L 1, and hybrid FIR state estimators. Such estimators are called robust, but their development for practical use has been carried out to a much lesser extent and awaits further development.

In this chapter, we provide examples of practical applications of FIR state estimators in various fields and different environments using batch and iterative algorithms.

11.1 UFIR Filtering and Prediction of Clock States

The IEEE Standard 1139‐2008 [75] and International Telecommunication Union (ITU‐T) recommendation G.810 [77] suggest that a clock has three states, namely, the time interval error (TIE), fractional frequency offset, and linear frequency drift rate. The problem with accurate estimation of clock state has a lot to do with the clock oscillator noise, which has a slow flicker (technical) Gaussian component having the PSD slope 1 slash normal f in the Fourier frequency normal f domain. Even modified for CPN [182], which has the PSD slope 1 slash normal f squared, the KF is unable to track accurately the 1 slash normal f behaviors. In GPS‐based timekeeping using one pulse per second (1PPS) signals, the problem is complicated with the uniformly distributed (non‐Gaussian) measurement noise and GPS time uncertainties caused by different satellites in a view. Therefore, the UFIR filter, which ignores zero mean noise, better fulfils the requirement of the IEEE Standard 1139‐2008 [75] that “an efficient and unbiased estimator is preferred” for clock applications [166].

11.1.1 Clock Model

In accordance with [75] and [77], the clock TIE alpha left-parenthesis t right-parenthesis caused by oscillator instabilities is modeled in continuous time t with the finite Taylor series as

where alpha 0 equals alpha left-parenthesis 0 right-parenthesis is the initial TIE (first state) at t equals 0, beta 0 equals beta left-parenthesis 0 right-parenthesis is the initial fractional frequency offset (second state), and gamma 0 equals gamma left-parenthesis 0 right-parenthesis is the initial linear frequency drift rate (third state). Noise w Subscript alpha Baseline left-parenthesis t right-parenthesis equals phi left-parenthesis t right-parenthesis slash 2 pi nu Subscript nom is completely defined by the clock oscillator random phase deviation component phi left-parenthesis t right-parenthesis and nominal frequency nu Subscript nom in Hz. The IEEE standard [75] states that w Subscript alpha Baseline left-parenthesis t right-parenthesis is affected by different kinds of independent phase and frequency fluctuations (white, flicker, and random walks), which PSD slopes can be measured in the Fourier frequency domain.

In state space, (11.1) is represented with the SDE

in which x left-parenthesis t right-parenthesis equals left-bracket alpha left-parenthesis t right-parenthesis beta left-parenthesis t right-parenthesis gamma left-parenthesis t right-parenthesis right-bracket Superscript upper T is the clock state vector and w left-parenthesis t right-parenthesis equals left-bracket w Subscript beta Baseline left-parenthesis t right-parenthesis w Subscript gamma Baseline left-parenthesis t right-parenthesis w Subscript ModifyingAbove gamma With dot Baseline left-parenthesis t right-parenthesis right-bracket Superscript upper T is the zero mean Gaussian noise vector having the covariance upper Q Subscript w Baseline left-parenthesis t comma theta right-parenthesis equals script upper E left-brace w left-parenthesis t right-parenthesis w Superscript upper T Baseline left-parenthesis theta right-parenthesis right-brace. In this vector, w Subscript beta Baseline left-parenthesis t right-parenthesis is the frequency noise, w Subscript gamma Baseline left-parenthesis t right-parenthesis is the linear frequency drift noise, and w Subscript ModifyingAbove gamma With dot Baseline left-parenthesis t right-parenthesis is noise in the first time derivative of the linear frequency drift. Note that w left-parenthesis t right-parenthesis is nonstationary on a long time scale, although it is typically assumed to be stationary on a short time. The clock state transition matrix upper A is specified by

(11.3)upper A equals Start 3 By 3 Matrix 1st Row 1st Column 0 2nd Column 1 3rd Column 0 2nd Row 1st Column 0 2nd Column 0 3rd Column 1 3rd Row 1st Column 0 2nd Column 0 3rd Column 0 EndMatrix period

In discrete time index k, model (11.2) is represented with

(11.4)x Subscript k Baseline equals upper F x Subscript k minus 1 Baseline plus w overbar Subscript k Baseline comma

where the clock matrix upper F delta-equals upper F left-parenthesis tau right-parenthesis is defined by the matrix exponential upper F equals e Superscript upper A tau to be

(11.5)upper F equals Start 3 By 3 Matrix 1st Row 1st Column 1 2nd Column tau 3rd Column StartFraction tau squared Over 2 EndFraction 2nd Row 1st Column 0 2nd Column 1 3rd Column tau 3rd Row 1st Column 0 2nd Column 0 3rd Column 1 EndMatrix period

The zero mean Gaussian noise vector w overbar Subscript k and its covariance upper Q Subscript w overbar Baseline left-parenthesis i comma j right-parenthesis equals script upper E left-brace w overbar Subscript i Baseline w overbar Subscript j Superscript upper T Baseline right-brace are defined by, respectively,

and we notice that (11.6) and (11.7) can be applied to all kinds of phase and frequency noise sources.

In timekeeping, only the first clock state (TIE) is commonly measured. Therefore, the observation equation becomes

(11.8)y Subscript k Baseline equals upper H x Subscript k Baseline plus v Subscript k Baseline comma

where upper H equals left-bracket 1 0 0 right-bracket, and v Subscript k is zero mean noise, which is not Gaussian in the 1PPS output of a GPS timing receiver.

11.1.2 Clock State Estimation Over GPS‐Based TIE Data

To compare errors produced by the KF and UFIR filter, the TIE measurements of a clock embedded in the Frequency Counter SR620 were provided using another SR620 as shown in [179]. The 1PPS output of the GPS SynPaQ III Timing Sensor was used as a reference signal. The ground truth was obtained for the Cesium Frequency Standard CsIII time signals.

UFIR Filtering Estimate

The iterative a posteriori UFIR filtering Algorithm 11 can now be applied to estimate the clock state. A specific feature is that due to the extremely slowly changing TIE of a precise clock and small noise, the optimal averaging horizon appears to be very large, upper N Subscript opt Baseline approximately-equals 3500 [168].

Kalman Filtering Estimate

To apply the a posteriori KF algorithm, it needs specifying correctly noise covariances and initial clock state. Noise in the clock oscillator embedded into SR620 is characterized with three values of the Allan deviation: sigma Subscript beta Baseline left-parenthesis 1 s right-parenthesis equals 2.3 times 1 0 Superscript negative 11, sigma Subscript beta Baseline left-parenthesis 10 s right-parenthesis equals 1.0 times 1 0 Superscript negative 11, and sigma Subscript beta Baseline left-parenthesis 100 s right-parenthesis equals 4.2 times 1 0 Superscript negative 11. Following [32], these values can be converted to the diffusion parameters q 1, q 2, and q 3 via

(11.9)sigma Subscript beta Superscript 2 Baseline left-parenthesis tau right-parenthesis equals StartFraction q 1 Over tau EndFraction plus StartFraction q 2 tau Over 3 EndFraction plus StartFraction q 3 tau cubed Over 20 EndFraction

and then the noise covariance upper Q Subscript w overbar Baseline left-parenthesis tau right-parenthesis specified in white Gaussian approximation as [192]

(11.10)StartFraction upper Q Subscript w overbar Baseline left-parenthesis tau right-parenthesis Over tau EndFraction equals Start 3 By 3 Matrix 1st Row 1st Column q 1 plus StartFraction q 2 tau squared Over 3 EndFraction plus StartFraction q 3 tau Superscript 4 Baseline Over 20 EndFraction 2nd Column StartFraction q 2 tau Over 2 EndFraction plus StartFraction q 3 tau cubed Over 8 EndFraction 3rd Column StartFraction q 3 tau squared Over 6 EndFraction 2nd Row 1st Column StartFraction q 2 tau Over 2 EndFraction plus StartFraction q 3 tau cubed Over 8 EndFraction 2nd Column q 2 plus StartFraction q 3 tau squared Over 3 EndFraction 3rd Column StartFraction q 3 tau Over 2 EndFraction 3rd Row 1st Column StartFraction q 3 tau squared Over 6 EndFraction 2nd Column StartFraction q 3 tau Over 2 EndFraction 3rd Column q 3 EndMatrix period

Note that since sigma Subscript normal y Superscript 2 Baseline left-parenthesis tau right-parenthesis is upper‐bounded in the clock oscillator specification, it can be reduced by the factor of 2 for the KF to perform better. Finally, the variance of the sawtooth noise induced by GPS SynPaQ III Timing Sensor is measured as upper Q Subscript v Baseline equals 5 0 squared slash 3 nssquared, and the unknown initial states can be set as alpha left-parenthesis 0 right-parenthesis equals alpha 0, beta left-parenthesis 0 right-parenthesis equals 0, and gamma left-parenthesis 0 right-parenthesis equals 0.

Typical estimates of the clock TIE alpha Subscript k and fractional frequency offset beta Subscript k are sketched in Fig. 11.1 along with the GPS‐based measurements of the TIE and ground truth provided by the Cesium Frequency Standard CsIII. The results reveal that efforts made to describe the clock noise covariance upper Q Subscript w overbar Baseline left-parenthesis tau right-parenthesis via the Allan deviation sigma Subscript beta Baseline left-parenthesis tau right-parenthesis for the KF were less successful than to measure experimentally upper N Subscript opt for the UFIR filter. Consequently, the KF produces the worst estimates even when the UFIR filter is tuned in a wide range of upper N, from 1500 to 3500. It also reveals that errors in the KF can be reduced by decreasing the Allan variance sigma Subscript beta Superscript 2 Baseline left-parenthesis tau right-parenthesis. But this takes sigma Subscript beta Superscript 2 Baseline left-parenthesis tau right-parenthesis values outside the scope of physical imagination and has no theoretical justification. It can also be seen that a near optimal horizon upper N Subscript opt Baseline equals 3500 makes the UFIR filter much more accurate than the KF. All that follows from this experiment is that the UFIR filter is more suitable for estimation of clock state than the KF.

Schematic illustration of typical estimates of the clock TIE produced by the UFIR filter and KF for the GPS-based TIE measurement and ground truth obtained by the Cesium Frequency Standard CsIII: (a) TIE and (b) fractional frequency offset.

Figure 11.1 Typical estimates of the clock TIE produced by the UFIR filter and KF for the GPS‐based TIE measurement and ground truth obtained by the Cesium Frequency Standard CsIII: (a) TIE and (b) fractional frequency offset.

It should be noted that the most accurate estimator for precise clocks is the batch OFIR filter, for which the experimentally measured full block covariance matrix script í’¬ Subscript upper N should contain the necessary information about the clock colored noise. This matrix can be large, 3500 times 3500 in the previous case, and it may take time to provide the estimation. But in most cases this can be tolerated due to the slow error processes in precise clocks.

11.1.3 Master Clock Error Prediction

Error prediction in national master clocks (MCs) is required, because the International Bureau of Weights and Measures (BIPM) determines time deviations with a five‐day interval as average values per day and issues the results monthly. As an example, we consider the UTC–UTC(NIST MC) time differences (285 points) measured each 10 days in 2002–2009 for the National Institute of Standards and Technology (NIST). The time scale is formed starting at k equals 0 [52279 Modified Julian Date (MJD)] and finishing at n equals 284 (55129 MJD) as shown in Fig. 11.2a. The MCs are commonly modeled with two states, upper K equals 2. Because the NIST MC is error‐corrected, we suppose that the process is stationary and employ all data available. Accordingly, we let upper N equals k plus 1 and use the full horizon p‐step UFIR predictor to bridge a month data gap. All along we also use the KF tuned in the best way.

Schematic illustration of estimates of the NIST MC current state via the UTC-UTC(NIST MC) time differences (285 points) measured in 2002-2009 each 10 days: (a) TIE prediction and (b) fractional frequency offset. Digits indicate the time points from which the arrows come out.

Figure 11.2 Estimates of the NIST MC current state via the UTC–UTC(NIST MC) time differences (285 points) measured in 2002–2009 each 10 days: (a) TIE prediction and (b) fractional frequency offset. Digits indicate the time points from which the arrows come out.

A highly useful property of the full horizon p‐step UFIR predictor is that it needs only a step p for interpolation or extrapolation. Because the NIST MC exhibits excellent etalon properties, the initial conditions can be set to zero, alpha 0 equals 0 and beta 0 equals 0. For the resolution of 0.1 ns in the published data, the uniformly distributed digitization noise has the variance sigma Subscript v Superscript 2 Baseline equals 0.0 5 squared slash 3 nssquared equals 8.33 times 1 0 Superscript negative 4 nssquared.

Figure 11.2 sketches estimates of alpha Subscript k and beta Subscript k produced by the UFIR filter and KF. Current estimates of the TIE alpha Subscript k are shown in Fig. 11.2a along with p‐step predicted values. Predictions are depicted with arrows coming out from the points indicated with digits. Since the UFIR estimator is unbiased, the first arrow coincides in the direction with several initial measurement points. With increased k, the prediction vectors show possible behaviors extrapolated at k plus p, p greater-than 0, over measurements from 0 to k. An estimator also reveals a small positive angle resulting in the frequency offset shown in Fig. 11.2b.

Analyzing Fig. 11.2, one may conclude that the KF is less suitable for MCs than the UFIR filter. Indeed, due to transients and a small number of data points, the KF does not sketch a real error picture at the initial stage. In the intermediate region (about the point of 1 times 1 0 cubed days), both filters produce consistent estimates. On a long time scale, the full horizon UFIR filter improves estimates, while the KF performs equally.

11.2 Suboptimal Clock Synchronization

The need to synchronize local time scales [76] arises with different allowed uncertainties in digital communications, bistatic radars, telephone networks, networked measurement and control systems, space systems, and computer nets. To discipline clocks, commercially available GPS timing receivers are often used to convey the reference time to the locked clock loop via the 1PPS output. The loop is organized for the clock TIE to range over time below an allowed threshold that, for digital communication networks, is specified in [78]. The GPS‐based clock steering is typically organized in two ways:

  • The frequency offset is adjusted in the clock oscillator, and the TIE is adjusted in the clock digital block.
  • In the clock digital block, only the TIE is adjusted if the oscillator is uncontrolled.

Many designs use the 1PPS output of a GPS timing receiver as a time reference. The time difference between the 1PPS signal, which is accurate but not precise, and the 1 normal s output of the local clock, which is precise but not accurate, is periodically measured using a high‐resolution TIE counter. The clock TIE is then estimated by a filter to obtain a synchronizing signal intended to discipline the local clock time scale as shown in [11, 170].

In locked clocks, the TIE noise is mainly dependent on the precision of a local oscillator, and the TIE departures are due to the limited accuracy of the reference time signals. The required synchronization accuracy can be achieved using various filters. Averaging and LP filters are typical for commercially manufactured locked clocks. In some designs, an integrating filter, linear LS estimator, KF, or even neural network is included in a disciplining phase‐locked loop (PLL) [11].

Schematic illustration of loop model of local clock synchronization based on GPS 1PPS timing signals.

Figure 11.3 Loop model of local clock synchronization based on GPS 1PPS timing signals.

11.2.1 Clock Digital Synchronization Loop

The clock synchronization loop proposed in [11] is shown in Fig. 11.3. The GPS 1PPS timing signal s Subscript k is represented with the GPS time uncertainty (or disturbance) w Subscript k caused by different satellites in a view and some other random factors. It is also complicated by the uniformly distributed sawtooth noise v Subscript k, whose values range from negative 50 ns to 50 ns owing to the principle of the 1PPS signal formation. Typical examples of the GPS 1PPS signal errors are given in Fig. 11.4 [11]. Random w Subscript k (Fig. 11.4a) ranges closer to zero than the sawtooth noise v Subscript k, which is uniformly distributed within left-bracket minus 50 ellipsis 50 right-bracket in ns (Fig. 11.4b). Errors in the GPS 1PPS timing signal shown in Fig. 11.4c are caused by a mixture of w Subscript k and v Subscript k. In some GPS timing receivers, a negative sawtooth correction code is available in the protocol. If this code is applied, the 1PPS signal error becomes approximately s Subscript n Baseline equals v Subscript n.

Schematic illustration of typical errors in GPS timing receivers: (a) GPS time uncertainty wk caused by different satellites in a view and other random factors, (b) sawtooth noise vk induced by the receiver, and (c) time error sk in the GPS 1PPS reference signal.

Figure 11.4 Typical errors in GPS timing receivers: (a) GPS time uncertainty w Subscript k caused by different satellites in a view and other random factors, (b) sawtooth noise v Subscript k induced by the receiver, and (c) time error s Subscript k in the GPS 1PPS reference signal.

Schematic illustration of a typical function of a nonstationary TIE αk of an unlocked crystal clock.

Figure 11.5 A typical function of a nonstationary TIE alpha Subscript k of an unlocked crystal clock.

A typical function of a nonstationary TIE alpha Subscript n of an unlocked crystal clock is shown in Fig. 11.5 [11]. The synchronization loop requires upper N neighboring past data points, from k minus upper N to k minus 1, and operates as follows. The TIE alpha Subscript k is measured relative to s Subscript k by the TIE counter, and the difference signal y Subscript k Baseline equals s Subscript k Baseline minus alpha Subscript k is computed. To predict the disciplining signal u Subscript k at the current point k, the RH UFIR filter is used, which has an impulse response of the ith degree h Subscript k Superscript left-parenthesis i right-parenthesis Baseline left-parenthesis 1 right-parenthesis. The predicted value u overtilde Subscript k is held as u overbar Subscript k by the hold filter, smoothed by the LP filter with the impulse response h Subscript LP k, and scaled with kappa to obtain the necessary control signal ModifyingAbove a With caret Subscript k, which disciplines the clock.

The ramp RH UFIR filter of the first degree, i equals 1 and p equals 1, is used with the impulse response function originally derived in [71],

(11.11)h Subscript k Superscript left-parenthesis 1 right-parenthesis Baseline left-parenthesis 1 right-parenthesis equals StartLayout Enlarged left-brace 1st Row 1st Column StartFraction 2 left-parenthesis 2 upper N plus 1 right-parenthesis minus 6 k Over upper N left-parenthesis upper N minus 1 right-parenthesis EndFraction 2nd Column comma 3rd Column 1 less-than-or-slanted-equals k less-than-or-slanted-equals upper N 2nd Row 1st Column 0 2nd Column comma 3rd Column otherwise EndLayout period

The 1‐step predictive filtering estimate is obtained as

(11.12)u overtilde Subscript k Baseline equals sigma-summation Underscript n equals 1 Overscript upper N Endscripts h Subscript n Superscript left-parenthesis 1 right-parenthesis Baseline left-parenthesis 1 right-parenthesis y Subscript k minus n

by the discrete convolution using data y Subscript k taken from left-bracket k minus upper N comma k minus 1 right-bracket.

To hold u overtilde Subscript k over time to continuously steer the clock between suboptimally defined values, a hold filter is used such that

where left floor StartFraction k Over upper M EndFraction right floor is an integer part of k slash upper M. By (11.13), the input and output values of the hold filter become equal when k is multiple to upper M. Between two such adjacent points, the output value of the hold filter is constant.

For multiple clock steering with period tau upper M, where tau is the sampling time and upper M is the number of sampling intervals, the hold filter produces a step signal u overbar Subscript k. Applied directly to the clock, u overbar Subscript k guarantees suboptimal steering and assures that alpha Subscript k ranges within narrow bounds around zero. On the other hand, the stepwise u overbar Subscript k may not guarantee the required clock performance.

To smooth u overbar Subscript k, an LP filter with the impulse response h Subscript LP k is used. It has been shown in [11] experimentally that the 1‐order LP filter with the impulse response

(11.14)h Subscript LP k Baseline equals StartLayout Enlarged left-brace 1st Row 1st Column a e Superscript minus StartFraction tau Over upper T EndFraction k Baseline 2nd Column comma 3rd Column k greater-than-or-equal-to 0 2nd Row 1st Column 0 2nd Column comma 3rd Column k less-than 0 EndLayout comma

where upper T is the filter time constant and a equals 1 minus e Superscript negative tau slash upper T, is able to fit the demands for precise clock synchronization. Finally, the gain kappa is applied to produce the control signal u Subscript k and provide steering of clock errors.

Experimental Verification

Experimental investigations of the synchronization loop shown in Fig. 11.3 were provided in [11]. To outline the capabilities of FIR state estimators in practical applications, next we present only the most typical characteristics of a steered clock with an oven‐controlled crystal oscillator (OCXO). The clock errors are presented in terms of the Allan deviation and precision time protocol (PTP) variance for tau equals 1 normal s, upper M equals upper N, and kappa equals 1. The optimal horizon upper N Subscript opt and optimal time constant upper T Subscript opt are measured experimentally by minimizing the MSE relative to the TIE of an OCXO‐based clock with and without the sawtooth. For GPS 1PPS signal with sawtooth, the optimal horizon was measured as upper N Subscript opt Baseline equals 250 and without the sawtooth as upper N Subscript opt Baseline equals 150. More details about this loop can be found in [11].

Schematic illustration of allan deviation of the GPS-locked OCXO-based clock for different T,s of the 1-order LP filter. GPS 1PPS signal has sawtooth and Nopt=250 [11].

Figure 11.6 Allan deviation of the GPS‐locked OCXO‐based clock for different upper T comma normal s of the 1‐order LP filter. GPS 1PPS signal has sawtooth and upper N Subscript opt Baseline equals 250 [11].

Allan Deviation

The Allan deviation of a locked OCXO‐based clock was investigated in [11] as a function of the LP filter time constant upper T. It follows from Fig. 11.6 that the Allan deviation of the 1PPS signal with sawtooth ranges higher than that of the OCXO‐based clock when the averaging time tau overbar is less than about 5000 sec. The suboptimal UFIR filter removes the sawtooth, improves the performance, and makes it such that the Allan deviation remains large only when tau overbar less-than 2300 normal s. Further improvement is achieved using an LP filter, which dramatically reduces the Allan deviation to the level of an OCXO‐based clock. Figure 11.6 suggests that upper T approximately-equals 2000 normal s can be accepted as near optimal. Indeed, with small tau overbar the Allan deviation value is due to the unlocked clock and with large tau overbar due to the UFIR filter (or sawtooth‐less measurements). The LP filter causes an excursion, who peak value reaches a minimum when upper T Subscript opt Baseline approximately-equals 2000 normal s, it grows and shifts to the right when upper T greater-than upper T Subscript opt, and it grows and shifts to the left if upper T less-than upper T Subscript opt.

Precision Time Protocol Variance

Another measure of clock errors is the PTV variance. It is worth noting as a matter of notation that by upper M equals upper N the square root of PTP variance sigma Subscript PTP Superscript 2 Baseline left-parenthesis tau overbar right-parenthesis is equal to the time deviation (TDEV) specified in [78]. The PTP variance relates to the Allan variance sigma Subscript beta Superscript 2 Baseline left-parenthesis tau overbar right-parenthesis as sigma Subscript PTP Superscript 2 Baseline left-parenthesis tau overbar right-parenthesis equals tau overbar squared sigma Subscript y Superscript 2 Baseline slash 3, and [76] states that this is the main measure of locked clock errors.

The improvement in PTP variance for a GPS 1PPS signal measured without a sawtooth is illustrated in Fig. 11.7, where the recommended boundary (dashed) is taken from [78].

Schematic illustration of PTP deviation of a GPS locked crystal clock for different T,s of the 1-order LP filter and Nopt=150 over GPS 1PPS signal without the sawtooth.

Figure 11.7 PTP deviation of a GPS locked crystal clock for different upper T comma normal s of the 1‐order LP filter and upper N Subscript opt Baseline equals 150 over GPS 1PPS signal without the sawtooth.

As can be seen, the PTP deviation of an OCXO‐based clock exceeds the recommended boundary when tau overbar greater-than 7000 normal s. The PTP deviation of the GPS 1PPS signal without the sawtooth ranges very close to the UFIR filter output, and it follows that the UFIR filter removes the sawtooth very efficiently. Both these measures range below the recommended boundary and approach it when tau overbar approximately-equals 100 normal s. The LP filter significantly improves the performance and reduces the clock PTP deviation by a factor of 10 or more over the recommended boundary. It also follows that a near optimal time constant can be accepted as upper T Subscript opt Baseline approximately-equals 1000 normal s.

In general, the fact remains: the iterative UFIR filtering Algorithm 11 is an efficient tool for GPS‐based clock steering, since the KF is not suitable for filtering flicker noise components.

Schematic illustration of 2-D schematic geometry of the mobile robot localization.

Figure 11.8 2‐D schematic geometry of the mobile robot localization.

11.3 Localization Over WSNs Using Particle/UFIR Filter

A well‐known disadvantage of the PF‐based localization of maneuvering objects is the need to generate a large number of particles in order to avoid divergence. The process requires resampling, which assumes that particles with higher weights (i.e., high likelihoods) will be statistically selected many times. This leads to a loss of diversity among the particles so that the resultant set will contain many repeated particles. The effect is known as sample impoverishment and usually occurs when the noise is not intensive or the number of particles is small.

To make the PF‐based localization more reliable by avoiding the divergence, the PF was combined in [143] with the UFIR filter. In the proposed hybrid PF/UFIR structure, the PF plays the role of the main filter, and the UFIR filter is used as the supporting filter. The PF estimates the state in normal situations, when there is sample impoverishment and no failures. Otherwise, when the PF fails, the UFIR filter restarts it.

This hybrid structure was used in [143] to provide indoor mobile robot localization using a wireless tag with a transmitter, four receivers, and a server computer as shown in Fig. 11.8. A wireless tag attached to the mobile robot transmits a signal to four receivers deployed at exactly known coordinates. The clocks of the receivers are synchronized using the synchronization line. Distances between the tag and the receivers are measured via time‐of‐arrival (TOA). The TOA data are transferred to the server computer to generate time‐difference‐of‐arrival (TDOA) measurements, which are represented by the following equation,

(11.15)Start 3 By 1 Matrix 1st Row z Subscript 1 comma k Baseline 2nd Row z Subscript 2 comma k Baseline 3rd Row z Subscript 3 comma k Baseline EndMatrix equals Start 3 By 1 Matrix 1st Row h Subscript 1 comma k Baseline 2nd Row h Subscript 2 comma k Baseline 3rd Row h Subscript 3 comma k Baseline EndMatrix equals StartFraction 1 Over c EndFraction Start 3 By 1 Matrix 1st Row d 1 minus d 2 2nd Row d 1 minus d 3 3rd Row d 1 minus d 4 EndMatrix comma

where z Subscript 1 comma k, z Subscript 2 comma k, and z Subscript 3 comma k are the TDOA data (in units of nanoseconds) and c is the speed of light. Here, d Subscript i, i element-of left-bracket 1 comma 4 right-bracket, is the ith distances between the mobile robot and the receivers. The distances are coupled with the robot local coordinates normal x Subscript k and normal y Subscript k and the ith receiver constant coordinates normal x Subscript i and normal y Subscript i by four nonlinear equations for i element-of left-bracket 1 comma 4 right-bracket,

(11.16)d Subscript i Baseline equals StartRoot left-parenthesis normal x Subscript k Baseline minus normal x Subscript i Baseline right-parenthesis squared plus left-parenthesis normal y Subscript k Baseline minus normal y Subscript i Baseline right-parenthesis squared EndRoot period

At the current point k, the mobile robot pose is given by the state vector x Subscript k Baseline equals left-bracket normal x Subscript k Baseline normal y Subscript k Baseline theta Subscript k Baseline right-bracket Superscript upper T, where theta Subscript k is a heading angle in the 2D plane local coordinates. Motion of the mobile robot is adjusted by the control signal u Subscript k Baseline equals left-bracket normal upper Delta d Subscript k Baseline normal upper Delta theta Subscript k Baseline right-bracket Superscript upper T, where normal upper Delta d Subscript k is the incremental distance (in meters) and normal upper Delta theta Subscript k is the incremental change in the heading angle (in degrees). The following difference equations represent the dynamics of the robot's movement,

(11.17)normal x Subscript k Baseline equals f Subscript 1 comma k Baseline equals normal x Subscript k minus 1 Baseline plus normal upper Delta d cosine left-parenthesis theta Subscript k minus 1 Baseline plus one half normal upper Delta theta Subscript k Baseline right-parenthesis comma
(11.18)normal y Subscript k Baseline equals f Subscript 2 comma k Baseline equals normal y Subscript k minus 1 Baseline plus normal upper Delta d sine left-parenthesis theta Subscript k minus 1 Baseline plus one half normal upper Delta theta Subscript k Baseline right-parenthesis comma
(11.19)theta Subscript k Baseline equals f Subscript 3 comma k Baseline equals theta Subscript k minus 1 Baseline plus normal upper Delta theta Subscript k Baseline period

The robot is equipped with a fiber optic gyroscope (FOG), which directly measures theta Subscript k, and the fourth required measurement becomes

(11.20)z Subscript 4 comma k Baseline equals h Subscript 4 comma k Baseline equals theta Subscript k Baseline period

Combined the measurements, the observation vector is constructed as y Subscript k Baseline equals left-bracket z Subscript 1 comma k Baseline z Subscript 2 comma k Baseline z Subscript 3 comma k Baseline z Subscript 4 comma k Baseline right-bracket Superscript upper T. Finally, the state equation f Subscript k Baseline equals left-bracket f Subscript 1 comma k Baseline f Subscript 2 comma k Baseline f Subscript 3 comma k Baseline right-bracket Superscript upper T and observation equation h Subscript k Baseline equals left-bracket h Subscript 1 comma k Baseline h Subscript 2 comma k Baseline h Subscript 3 comma k Baseline h Subscript 4 comma k Baseline right-bracket Superscript upper T formalize the localization problem in state space with

where the process noise w Subscript k Baseline tilde script í’© left-parenthesis 0 comma upper Q Subscript k Baseline right-parenthesis and measurement noise v Subscript k Baseline tilde script í’© left-parenthesis 0 comma upper R Subscript k Baseline right-parenthesis have the covariances upper Q Subscript k and upper R Subscript k. Given (11.21) and (11.22), the mobile robot coordinates and heading are estimated by the PF.

11.3.1 Sample Impoverishment Issue

The PF‐based approach assumes that the robot coordinates are the first‐order Markov processes evolving from one point to another with known initial and transition distributions and that the conditionally independent observations depend only on the robot position. To estimate normal x Subscript k, normal y Subscript k, and theta Subscript k, the PF generates a set of samples at each k that approximates the distributions of the coordinates conditioned on all past observations. This process is called resampling and causes an issue known as sample impoverishment.

Schematic illustration of a typical scenario with the sample impoverishment.

Figure 11.9 A typical scenario with the sample impoverishment.

Figure 11.9 illustrates an idea of sample impoverishment in a typical scenario. The ith dot, i element-of left-bracket 1 comma upper L right-bracket, represents a sample of the ith predicted measurement y Subscript k comma i Superscript minus defined by

(11.23)y Subscript k comma i Superscript minus Baseline equals h Subscript k Baseline left-parenthesis ModifyingAbove x With caret Subscript k comma i Superscript minus Baseline right-parenthesis comma

where ModifyingAbove x With caret Subscript k comma i Superscript minus is the ith a priori particle (i.e., sample of the a priori estimated state) and upper L is the number of the particles. In the Gaussian approximation, the likelihood (i.e., weight) of each particle is a reciprocal of the difference between the actual measurement y Subscript k Superscript asterisk and the predicted measurement. Therefore, the closer the samples of the predicted measurement y Subscript k comma i Superscript minus are to y Subscript k Superscript asterisk, the higher the weights of the corresponding a priori particles.

In the example shown in Fig. 11.9, there are only two samples of predicted measurements within the measurement: uncertainty or error ellipse. Thus, only these particles will receive significant weights, and resampling will repeat them many times as the a posteriori particles ModifyingAbove x With caret Subscript k comma i. As a consequence, sample impoverishment and failures will occur. The problem is mainly associated with low‐intensity process noise and/or measurement noise and/or the small number of particles required to ensure fast localization. Other factors can also cause sample impoverishment. Therefore, a great deal of efforts may be required to satisfactorily address the problem. But despite all efforts, the problem remains fundamental, and it is impossible to completely avoid sample impoverishment.

11.3.2 Hybrid Particle/UFIR Filter

To improve localization accuracy, PF has been combined in [143] with an EFIR filter in a hybrid algorithm that encompasses two key procedures: 1) detection of PF failures and 2) particles regeneration and PF resetting. PF fault detection is organized in a diagnosis algorithm using Mahalanobis distance. To regenerate new particles and reset PF, an EFIR filter is used to obtain the state estimates when the PF fails. The choice of an EFIR filter that is robust and BIBO stable was made in attempts to fulfil a basic requirement: an auxiliary filter does not have to be obligatorily accurate, but it has to be robust and stable. Note that resetting the PF with an EFIR filter instead of generating a lot particles solves another critical issue: reducing the computational complexity and time required for real‐time localization.

Schematic illustration of a flowchart of the hybrid PF/EFIR algorithm.

Figure 11.10 A flowchart of the hybrid PF/EFIR algorithm.

A flowchart of the hybrid PF/EFIR algorithm is given in Fig. 11.10. PF plays the role of the main filter, which provides estimates of the mobile robot coordinates and heading under normal conditions. Diagnostics of PF failures is performed continuously, and when the PF fails, the auxiliary EFIR filter provides information to reset the failed PF. Erroneous estimates at the PF output are detected taking into account the following features of sample impoverishment: 1) only a few predicted samples fall into the uncertainty ellipse, and 2) the predicted samples are usually far from actual measurement. PF rejections are detected by checking the number of samples within the uncertainty ellipse using the Mahalanobis distance.

This example provides further evidence of the effectiveness of hybrid filters in solving the localization problem.

11.4 Self‐Localization Over RFID Tag Grids

Radio frequency identification (RFID) tag‐based networks and grids are designed to organize self‐localization of various moving objects in GPS‐denied indoor environments. Each tag has an ID number corresponding to a unique location and can be either active or passive. To increase awareness, information describing a local 2D or 3D surrounding can be programmed in each tag and delivered to users by request. The method is low cost and available for any purpose, provided the communication between a target and the tags. However, since low‐cost measurements are usually very noisy, optimal estimators are required.

Schematic illustration of 2D schematic geometry of a vehicle traveling on an indoor floorspace nested with two RFID tags, T1 and T2, having exactly known coordinates.

Figure 11.12 2D schematic geometry of a vehicle traveling on an indoor floorspace nested with two RFID tags, T1 and T2, having exactly known coordinates.

An example of the 2D schematic geometry of a vehicle traveling on an indoor floorspace nested with two RFID tags normal upper T 1 and normal upper T 2 (the number can be arbitrary) having exactly known coordinates is shown in Fig. 11.12 [149,150]. The vehicle reader measures the distances to the tags, and the heading angle upper Phi is measured using a fiber‐optic gyroscope (FOG); subfigures (a) and (b) are used when the vehicle and the tags are not in the same plane. The vehicle travels in direction d, and its trajectory is controlled by the left and right wheels. The incremental distances that the vehicle travels on these wheels are d Subscript normal upper L and d Subscript normal upper R, respectively. The distance between the left and right wheels is b, and the stabilized wheel is not shown. The vehicle moves in its own planar Cartesian coordinates left-parenthesis normal x Subscript normal r Baseline comma normal y Subscript normal r Baseline right-parenthesis with a center at upper M left-parenthesis normal x comma normal y right-parenthesis; that is, the vehicle direction always coincides with axis normal x Subscript normal r. The FOG measures upper Phi directly.

The indoor space is commonly nested with upper L RFID tags normal upper T Subscript l Baseline left-parenthesis chi Subscript l Baseline comma mu Subscript l Baseline right-parenthesis, l element-of left-bracket 1 comma upper L right-bracket, where coordinates left-parenthesis chi Subscript l Baseline comma mu Subscript l Baseline right-parenthesis of the lth tag are exactly known. It is supposed that at each k the vehicle reader can simultaneously detect kappa Subscript k Baseline greater-than-or-equal-to 2 tags, where the number kappa Subscript k varies over time, and measures the distance d Subscript i to the ith tag, where i element-of left-bracket 1 comma kappa Subscript k Baseline right-bracket, which can be any of the nested tags.

Referring to Fig. 11.12 and vehicle odometry, the incremental distance d Subscript k and heading angle phi Subscript k can be found as

(11.25)d Subscript k Baseline equals one half left-parenthesis d Subscript normal upper R k Baseline plus d Subscript normal upper L k Baseline right-parenthesis comma
(11.26)phi Subscript k Baseline approximately-equals StartFraction 1 Over b EndFraction left-parenthesis d Subscript normal upper R k Baseline minus d Subscript normal upper L k Baseline right-parenthesis

and the vehicle coordinates normal x Subscript k and normal y Subscript k and heading upper Phi Subscript k obtained by the vehicle kinematics using equations

(11.28)f Subscript 2 k Baseline equals normal y Subscript n Baseline equals normal y Subscript k minus 1 Baseline plus d Subscript k Baseline sine left-parenthesis upper Phi Subscript k minus 1 Baseline plus StartFraction phi Subscript k Baseline Over 2 EndFraction right-parenthesis comma

where normal x Subscript k minus 1, normal y Subscript k minus 1, and upper Phi Subscript k minus 1 are projected to k via the incremental distances d Subscript normal upper L k and d Subscript normal upper R k. Note that all these values are practically not exact and undergo random variations.

11.4.1 State‐Space Localization Problem

To solve the localization problem in state space, the state vector can be assigned as x Subscript k Baseline equals left-bracket normal x Subscript k Baseline normal y Subscript k Baseline upper Phi Subscript k Baseline right-bracket Superscript upper T and the input vector as u Subscript k Baseline equals left-bracket d Subscript normal upper L k Baseline d Subscript normal upper R k Baseline right-bracket Superscript upper T. Random additive components in these vectors can be supposed to be w Subscript k Baseline equals left-bracket w Subscript normal x k Baseline w Subscript normal y k Baseline w Subscript upper Phi k Baseline right-bracket Superscript upper T Baseline tilde script í’© left-parenthesis 0 comma upper Q right-parenthesis and e Subscript k Baseline equals left-bracket e Subscript normal upper L k Baseline e Subscript normal upper R k Baseline right-bracket Superscript upper T Baseline tilde script í’© left-parenthesis 0 comma upper L right-parenthesis. The nonlinear state equation thus becomes

(11.30)x Subscript k Baseline equals f Subscript k Baseline left-parenthesis x Subscript k minus 1 Baseline comma u Subscript k Baseline comma w Subscript k Baseline comma e Subscript k Baseline right-parenthesis comma

where the nonlinear function f Subscript k Baseline equals left-bracket f Subscript 1 k Baseline f Subscript 2 k Baseline f Subscript 3 k Baseline right-bracket Superscript upper T is combined with components given by (11.27)(11.29).

The measurement equations can be written as

StartLayout 1st Row 1st Column d Subscript 1 k 2nd Column equals 3rd Column StartRoot left-parenthesis mu 1 minus normal y Subscript k Baseline right-parenthesis squared plus left-parenthesis chi 1 minus normal x Subscript k Baseline right-parenthesis squared plus c 1 squared EndRoot comma 2nd Row 1st Column vertical-ellipsis 2nd Column Blank 3rd Row 1st Column d Subscript kappa Sub Subscript k Subscript k 2nd Column equals 3rd Column StartRoot left-parenthesis mu Subscript kappa Sub Subscript k Subscript Baseline minus normal y Subscript k Baseline right-parenthesis squared plus left-parenthesis chi Subscript kappa Sub Subscript k Subscript Baseline minus normal x Subscript n Baseline right-parenthesis squared plus c Subscript kappa Sub Subscript k Subscript Superscript 2 Baseline EndRoot comma 4th Row 1st Column upper Phi Subscript k 2nd Column equals 3rd Column upper Phi Subscript k Baseline period EndLayout

By introducing the observation vector y Subscript k Baseline equals left-bracket z Subscript 1 k Baseline ellipsis y Subscript kappa Sub Subscript k Subscript k Baseline y Subscript phi k Baseline right-bracket Superscript upper T Baseline element-of double-struck upper R Superscript kappa Super Subscript k Superscript plus 1, nonlinear function h Subscript k Baseline left-parenthesis x Subscript k Baseline right-parenthesis equals left-bracket d Subscript 1 k Baseline ellipsis d Subscript kappa Sub Subscript k Subscript k Baseline upper Phi Subscript k Baseline right-bracket Superscript upper T Baseline element-of double-struck upper R Superscript kappa Super Subscript k Superscript plus 1, and measurement noise v Subscript k Baseline equals left-bracket v Subscript 1 k Baseline ellipsis v Subscript kappa Sub Subscript k Subscript k Baseline v Subscript phi k Baseline right-bracket element-of double-struck upper R Superscript kappa Super Subscript k Superscript plus 1, the observation equation can be rewritten in a compact form as

(11.31)y Subscript k Baseline equals h Subscript k Baseline left-parenthesis x Subscript k Baseline right-parenthesis plus v Subscript k Baseline comma

where noise v Subscript k Baseline tilde script í’© left-parenthesis 0 comma upper R Subscript k Baseline right-parenthesis has the covariance upper R Subscript k Baseline equals script upper E left-brace v Subscript k Baseline v Subscript k Superscript upper T Baseline right-brace.

Extended State‐Space Model

The standard procedure (3.228) applied to function f Subscript k Baseline delta-equals f Subscript k Baseline left-parenthesis x Subscript k minus 1 Baseline comma u Subscript k Baseline comma w Subscript k Baseline comma e Subscript k Baseline right-parenthesis gives the first‐order approximation

(11.32)StartLayout 1st Row 1st Column f Subscript k 2nd Column approximately-equals 3rd Column f Subscript k Baseline left-parenthesis ModifyingAbove x With caret Subscript k minus 1 Baseline comma u Subscript k Baseline comma 0 comma 0 right-parenthesis plus upper F Subscript k Baseline left-parenthesis x Subscript k minus 1 Baseline minus ModifyingAbove x With caret Subscript k minus 1 Baseline right-parenthesis plus upper E Subscript k Baseline e Subscript k plus upper B Subscript k Baseline w Subscript k 2nd Row 1st Column Blank 2nd Column equals 3rd Column upper F Subscript k Baseline x Subscript k minus 1 Baseline plus u overbar Subscript k Baseline plus upper E Subscript k Baseline e Subscript k Baseline plus upper B Subscript k Baseline w Subscript k Baseline comma EndLayout

where ModifyingAbove x With caret Subscript k minus 1 is an estimate at k minus 1, u overbar Subscript k Baseline equals f Subscript k Baseline left-parenthesis ModifyingAbove x With caret Subscript k minus 1 Baseline comma u Subscript k Baseline comma 0 comma 0 right-parenthesis minus upper F Subscript k Baseline ModifyingAbove x With caret Subscript k minus 1 is known, and upper F Subscript k, upper B Subscript k, and upper E Subscript k are Jacobian matrices defined by

(11.33)upper F Subscript k Baseline equals StartFraction partial-differential f Subscript k Baseline Over partial-differential x EndFraction vertical-bar Subscript ModifyingAbove x With caret Sub Subscript k minus 1 Subscript Baseline equals Start 3 By 3 Matrix 1st Row 1st Column 1 2nd Column 0 3rd Column minus d Subscript k Baseline sine left-parenthesis ModifyingAbove upper Phi With Ì‚ Subscript k minus 1 Baseline plus one half phi Subscript k Baseline right-parenthesis 2nd Row 1st Column 0 2nd Column 1 3rd Column d Subscript k Baseline cosine left-parenthesis ModifyingAbove upper Phi With Ì‚ Subscript k minus 1 Baseline plus one half phi Subscript k Baseline right-parenthesis 3rd Row 1st Column 0 2nd Column 0 3rd Column 1 EndMatrix comma
(11.34)upper B Subscript k Baseline equals StartFraction partial-differential f Subscript k Baseline Over partial-differential w EndFraction vertical-bar Subscript ModifyingAbove x With caret Sub Subscript k minus 1 Subscript Baseline equals upper F Subscript k Baseline comma
(11.35)upper E Subscript k Baseline equals StartFraction 1 Over 2 b EndFraction Start 3 By 2 Matrix 1st Row 1st Column b e Subscript normal c k Baseline plus d Subscript k Baseline e Subscript normal s k Baseline 2nd Column b e Subscript normal c k Baseline minus d Subscript k Baseline e Subscript normal s k Baseline 2nd Row 1st Column b e Subscript normal s k Baseline minus d Subscript k Baseline e Subscript normal c k Baseline 2nd Column b e Subscript normal s k Baseline plus d Subscript n Baseline e Subscript normal c k Baseline 3rd Row 1st Column negative 2 2nd Column 2 EndMatrix comma

where e Subscript normal c k Baseline equals cosine left-parenthesis ModifyingAbove upper Phi With Ì‚ Subscript k Superscript minus Baseline plus StartFraction phi Subscript k Baseline Over 2 EndFraction right-parenthesis and e Subscript normal s k Baseline equals sine left-parenthesis ModifyingAbove upper Phi With Ì‚ Subscript k Superscript minus Baseline plus StartFraction phi Subscript k Baseline Over 2 EndFraction right-parenthesis.

By applying (3.229) to h Subscript k Baseline left-parenthesis x Subscript k Baseline right-parenthesis, the first‐order approximation becomes

(11.36)StartLayout 1st Row 1st Column h Subscript k Baseline left-parenthesis x Subscript k Baseline right-parenthesis 2nd Column approximately-equals 3rd Column h Subscript k Baseline left-parenthesis ModifyingAbove x With caret Subscript k Superscript minus Baseline right-parenthesis plus StartFraction partial-differential h Subscript k Baseline Over partial-differential x EndFraction vertical-bar Subscript ModifyingAbove x With caret Sub Subscript k Sub Superscript minus Baseline left-parenthesis x Subscript k Baseline minus ModifyingAbove x With caret Subscript k Superscript minus Baseline right-parenthesis 2nd Row 1st Column Blank 2nd Column equals 3rd Column upper H Subscript k Baseline x Subscript k Baseline plus y overbar Subscript k Baseline comma EndLayout

where y overbar Subscript k Baseline equals h Subscript k Baseline left-parenthesis ModifyingAbove x With caret Subscript k Superscript minus Baseline right-parenthesis minus upper H Subscript k Baseline ModifyingAbove x With caret Subscript k Superscript minus is known and upper H Subscript k is a Jacobian matrix defined by

(11.37)upper H Subscript k Baseline equals StartFraction partial-differential h Subscript k Baseline Over partial-differential x EndFraction vertical-bar Subscript ModifyingAbove x With caret Sub Subscript k Sub Superscript minus Subscript Baseline equals Start 4 By 3 Matrix 1st Row 1st Column StartFraction ModifyingAbove normal x With Ì‚ Subscript k Superscript minus Baseline minus chi overbar Subscript 1 Baseline Over nu Subscript 1 k Baseline EndFraction 2nd Column StartFraction ModifyingAbove normal y With Ì‚ Subscript k Superscript minus Baseline minus mu overbar Subscript 1 Baseline Over nu Subscript 1 k Baseline EndFraction 3rd Column 0 2nd Row 1st Column vertical-ellipsis 2nd Column vertical-ellipsis 3rd Column vertical-ellipsis 3rd Row 1st Column StartFraction ModifyingAbove normal x With Ì‚ Subscript k Superscript minus Baseline minus chi overbar Subscript kappa Sub Subscript k Subscript Baseline Over nu Subscript left-parenthesis kappa Sub Subscript k Subscript k right-parenthesis Baseline EndFraction 2nd Column StartFraction ModifyingAbove normal y With Ì‚ Subscript k Superscript minus Baseline minus mu overbar Subscript kappa Sub Subscript k Subscript Baseline Over nu Subscript left-parenthesis kappa Sub Subscript k Subscript k right-parenthesis Baseline EndFraction 3rd Column 0 4th Row 1st Column 0 2nd Column 0 3rd Column 1 EndMatrix comma

where nu Subscript i k Baseline equals StartRoot left-parenthesis mu overbar Subscript i Baseline minus ModifyingAbove normal y With Ì‚ Subscript k Superscript minus Baseline right-parenthesis squared plus left-parenthesis chi overbar Subscript i Baseline minus ModifyingAbove normal x With Ì‚ Subscript k Superscript minus Baseline right-parenthesis squared plus c Subscript i Superscript 2 Baseline EndRoot, i element-of left-bracket 1 comma kappa Subscript k Baseline right-bracket.

The first‐order extended state‐space model can therefore be written as

where w overTilde Subscript k Baseline tilde script í’© left-parenthesis 0 comma upper Q overTilde Subscript k Baseline right-parenthesis and e overTilde Subscript k Baseline tilde script í’© left-parenthesis 0 comma upper L overTilde Subscript k Baseline right-parenthesis have the covariances upper Q overTilde Subscript k Baseline equals upper F Subscript k Baseline upper Q upper F Subscript k Superscript upper T and upper L overTilde Subscript k Baseline equals upper E Subscript k Baseline upper L upper E Subscript k Superscript upper T.

11.4.2 Localization Performance

Provided the approximation (11.38) and (11.39), the EKF algorithm (3.240)–(3.244) and EFIR Algorithm 15 can be directly applied to organize self‐localization as shown next.

Consider a robot platform moving down an aisle in an RFID grid environment where tags are nested in the floor and ceiling (Fig. 11.13) [150].

Schematic illustration of schematic diagram of a vehicle platform traveling on an indoor passway in the RFID tag environment with eight tags mounted on a ceiling and eight tags mounted on a floor.

Figure 11.13 Schematic diagram of a vehicle platform traveling on an indoor passway in the RFID tag environment with eight tags mounted on a ceiling and eight tags mounted on a floor.

The environment is simple because the coordinates of a tag can be easily calculated based on its geometrical position. It is supposed that the reader is able to detect eight tags at the same time at each floorspace point, but some tags may not be available due to faults, furniture isolation, low‐power consumption, etc. However, at least two tags are always available.

To assess the quality of localization, noise was generated in [150] with the following standard deviations: sigma Subscript normal x Baseline equals sigma Subscript normal y Baseline equals 1 mm, sigma Subscript upper L Baseline equals sigma Subscript upper R Baseline equals 0.1 mm, and sigma Subscript upper Phi Baseline equals 0 period 5 Superscript ring. Since the tags on the ceiling are farthest from the platform, the standard deviations of the measurement noise are set as sigma Subscript v Baseline 1 Baseline equals midline-horizontal-ellipsis equals sigma Subscript v Baseline 8 Baseline equals 5 mm, sigma Subscript v Baseline 9 Baseline equals midline-horizontal-ellipsis equals sigma Subscript v Baseline 16 Baseline equals 10 mm, and sigma Subscript phi Baseline equals 2 Superscript ring. The optimal horizon is measured as upper N Subscript opt Baseline equals 12. The tags detected in 6 intervals (in m) along the passway are listed in Table 11.1, where tags 12 and 14–16 are not available.

Table 11.1 Tags detected in six intervals (in m) along a passway shown in Fig. 11.13; tags 12 and 14–16 were not available.

mTag
1  2  3  4  5  6  7  8  9  10  11  13
0–2xxxx
2–4xxxxx
4–6xxx
6–8xxxx
8–10xxx
10–12xx

Since noise statistics are typically not known exactly, an error factor p greater-than 0 was introduced such that p equals 1 makes the case ideal and the noise covariances were replaced in the algorithms by p squared upper R Subscript k, upper Q slash p squared, and upper L slash p squared. It has been shown that under ideal conditions, p equals 1, the EKF and EFIR filter give very consistent estimates and are not prone to divergence. The situation changes dramatically for p less-than 0.5, which leads to the errors shown in Fig. 11.14. Indeed, for tag sets taken from Table 11.1 with three different realisations of the same noise, EKF demonstrates: (a) local instability, (b) single divergence, and (c) multiple divergence. In contrast, the EFIR filter turned out to be much more stable with no signs of divergence.

Schematic illustration of localization errors caused by imprecisely known noise covariances for pltltlt0.5: (a) local instabilities, (b) single divergence of EKF, and (c) multiple divergence of EKF.

Figure 11.14 Localization errors caused by imprecisely known noise covariances for p less-than 0.5: (a) local instabilities, (b) single divergence of EKF, and (c) multiple divergence of EKF.

This investigation clearly points to another possible source of divergence in the EKF, namely, errors in the noise covariances. It also shows that increasing the number of tags interacting with the platform protects EKF from divergency so that massively nested tags can guarantee localization stability.

11.5 INS/UWB‐Based Quadrotor Localization

One of the most useful modern innovations is an unmanned aerial vehicle called a quadcopter or quadrotor. Due to the ability to fulfil tasks traditionally assigned to humans, quadrotors have became irreplaceable in many branches of modern life. A prerequisite for a quadrotor to complete various tasks is accurate position information, which must be retrieved quickly and preferably optimally with sufficient robustness in complex, harsh, and volatile navigation environments.

A typical quadrotor localization scheme that combines the capabilities of an inertial navigation system (INS) and ultra wide band (UWB)–based system is shown in Fig. 11.15 [217]. A navigation environment has been created here with multiple reference nodes (RNs), and the UWB and INS units are installed on the quadrotor. The UWB‐based subsystem derives the quadrotor position upper L Subscript k Superscript normal upper U Baseline equals left-parenthesis normal x Subscript k Superscript normal upper U Baseline comma normal y Subscript k Superscript normal upper U Baseline comma normal z Subscript k Superscript normal upper U Baseline right-parenthesis in the east direction normal x Subscript k Superscript normal upper U, north direction normal y Subscript k Superscript normal upper U, and vertical direction normal z Subscript k Superscript normal upper U via distances d Subscript i k Superscript normal upper U measured between the ith RN and a quadrotor. The INS‐based subsystem obtains the quadrotor position upper L Subscript k Superscript normal upper I Baseline equals left-parenthesis normal x Subscript k Superscript normal upper I Baseline comma normal y Subscript k Superscript normal upper I Baseline comma normal z Subscript k Superscript normal upper I Baseline right-parenthesis. Provided upper L Subscript k Superscript normal upper U and upper L Subscript k Superscript normal upper I, the difference delta upper L Subscript k Baseline equals upper L Subscript k Superscript normal upper I Baseline minus upper L Subscript k Superscript normal upper U is estimated to correct the INS position upper L Subscript k Superscript normal upper I and produce finally the position vector upper L Subscript k Baseline equals left-parenthesis normal x Subscript k Baseline comma normal y Subscript k Baseline comma normal z Subscript k Baseline right-parenthesis.

Schematic illustration of iNS/UWB-integrated quadrotor localization scheme [217].

Figure 11.15 INS/UWB‐integrated quadrotor localization scheme [217].

A specific feature of this system is that UWB data are contaminated with CMN in all of the directions as shown in Fig. 11.16. Thus, the white Gaussian approximation may not be sufficient for accurate quadrotor localization, and the GKF algorithm (3.146)–(3.151) or UFIR filtering Algorithm 13 should be applied.

Schematic illustration of CMN in UWB-derived data: (a) east direction, (b) north direction, and (c) vertical direction.

Figure 11.16 CMN in UWB‐derived data: (a) east direction, (b) north direction, and (c) vertical direction.

11.5.1 Quadrotor State Space Model Under CMN

Based on the scheme shown in Fig. 11.15, the quadrotor dynamics can be represented with increments in the coordinates delta normal x Subscript k, delta normal y Subscript k, and delta normal z Subscript k and velocities delta upper V Subscript normal x k, delta upper V Subscript normal y k, and delta upper V Subscript normal z k, which can then be united in the state equation

(11.40)x Subscript k Baseline equals upper F x Subscript k minus 1 Baseline plus w Subscript k Baseline comma

where

x Subscript k Baseline equals Start 6 By 1 Matrix 1st Row delta normal x Subscript k Baseline 2nd Row delta upper V Subscript normal x k Baseline 3rd Row delta normal y Subscript k Baseline 4th Row delta upper V Subscript normal y k Baseline 5th Row delta normal z Subscript k Baseline 6th Row delta upper V Subscript normal z k Baseline EndMatrix comma upper F equals Start 6 By 6 Matrix 1st Row 1st Column 1 2nd Column tau 3rd Column 0 4th Column 0 5th Column 0 6th Column 0 2nd Row 1st Column 0 2nd Column 1 3rd Column 0 4th Column 0 5th Column 0 6th Column 0 3rd Row 1st Column 0 2nd Column 0 3rd Column 1 4th Column tau 5th Column 0 6th Column 0 4th Row 1st Column 0 2nd Column 0 3rd Column 0 4th Column 1 5th Column 0 6th Column 0 5th Row 1st Column 0 2nd Column 0 3rd Column 0 4th Column 0 5th Column 1 6th Column tau 6th Row 1st Column 0 2nd Column 0 3rd Column 0 4th Column 0 5th Column 0 6th Column 1 EndMatrix comma

tau is the time step, and w Subscript k Baseline tilde script í’© left-parenthesis 0 comma upper Q right-parenthesis.

For delta upper L Subscript k Baseline equals upper L Subscript k Superscript normal upper I Baseline minus upper L Subscript k Superscript normal upper U representing measurements, the observation equation becomes

(11.41)z Subscript k Baseline equals upper H x Subscript k Baseline plus v Subscript k Baseline comma

where

z Subscript k Baseline equals Start 3 By 1 Matrix 1st Row delta normal x Subscript k Baseline 2nd Row delta normal y Subscript k Baseline 3rd Row delta normal z Subscript k Baseline EndMatrix comma upper H equals Start 3 By 6 Matrix 1st Row 1st Column 1 2nd Column 0 3rd Column 0 4th Column 0 5th Column 0 6th Column 0 2nd Row 1st Column 0 2nd Column 0 3rd Column 1 4th Column 0 5th Column 0 6th Column 0 3rd Row 1st Column 0 2nd Column 0 3rd Column 0 4th Column 0 5th Column 1 6th Column 0 EndMatrix comma

where v Subscript k Baseline equals left-bracket v Subscript normal x k Baseline v Subscript normal y k Baseline v Subscript normal z k Baseline right-bracket Superscript upper T is the CMN (Fig. 11.16) represented with the Gauss‐Markov model

(11.42)v Subscript k Baseline equals upper Psi v Subscript k minus 1 Baseline plus xi Subscript k Baseline comma

in which upper Psi is the coloredness factor matrix such that v Subscript k remains stationary and noise xi Subscript k Baseline tilde script í’© left-parenthesis 0 comma upper R Subscript xi Baseline right-parenthesis has the covariance upper R Subscript xi.

To apply the GKF and general UFIR filter, the measurement difference is introduced and transformed as

(11.43)StartLayout 1st Row 1st Column y Subscript k 2nd Column equals 3rd Column z Subscript k Baseline minus upper Psi z Subscript k minus 1 2nd Row 1st Column equals 2nd Column upper H Subscript k Baseline plus v Subscript k minus upper Psi upper H x Subscript k minus 1 minus upper Psi v Subscript k minus 1 3rd Row 1st Column Blank 2nd Column equals 3rd Column left-parenthesis upper H minus upper Xi right-parenthesis x Subscript k plus upper Xi w Subscript k plus xi Subscript k 4th Row 1st Column Blank 2nd Column equals 3rd Column upper C x Subscript k Baseline plus v overbar Subscript k Baseline comma EndLayout

where upper Xi equals upper Psi upper H upper F Superscript negative 1, upper C equals upper H minus upper Xi, and white Gaussian v overbar Subscript k Baseline equals upper Xi w Subscript k Baseline plus xi Subscript k Baseline tilde script í’© left-parenthesis 0 comma upper R overbar right-parenthesis has the the covariance

(11.44)upper R overbar equals script upper E left-brace v overbar Subscript k Baseline v overbar Subscript k Superscript upper T Baseline right-brace equals upper Xi upper Omega plus upper R comma

where upper Omega equals upper Q upper Xi Superscript upper T. For this model, the measurement residual s Subscript k is given by

(11.45)StartLayout 1st Row 1st Column s Subscript k 2nd Column equals 3rd Column y Subscript k Baseline minus upper C ModifyingAbove x With caret Subscript k 2nd Row 1st Column Blank 2nd Column equals 3rd Column upper C x Subscript k plus v overbar Subscript k minus upper C upper A ModifyingAbove x With caret Subscript k minus 1 3rd Row 1st Column Blank 2nd Column equals 3rd Column upper C upper A normal upper Delta Subscript k minus 1 Baseline plus upper C w Subscript k Baseline plus v overbar Subscript k Baseline comma EndLayout

where normal upper Delta Subscript k Baseline equals x Subscript k Baseline minus ModifyingAbove x With caret Subscript k. Since the cross covariance transforms to

script upper E left-brace v overbar Subscript k Baseline w Subscript k Superscript upper T Baseline right-brace equals script upper E left-brace left-parenthesis upper Xi w Subscript k Baseline plus xi Subscript k Baseline right-parenthesis w Subscript k Superscript upper T Baseline right-brace equals upper Xi upper Q

and upper P Subscript k Superscript minus Baseline equals upper A upper P Subscript k minus 1 Baseline upper A Superscript upper T Baseline plus upper Q is the prior error covariance, the innovation error covariance upper S Subscript k Baseline equals script upper E left-brace s Subscript upper K Baseline s Subscript k Superscript upper T Baseline right-brace becomes

(11.46)StartLayout 1st Row 1st Column upper S Subscript k 2nd Column equals 3rd Column upper C upper A upper P Subscript k minus 1 Baseline upper A Superscript upper T Baseline upper C Superscript upper T plus upper C upper Q upper C Superscript upper T plus upper Xi upper Omega plus upper R plus upper C upper Omega plus upper Omega Superscript upper T Baseline upper C Superscript upper T 2nd Row 1st Column Blank 2nd Column equals 3rd Column upper C upper P Subscript n Superscript minus Baseline upper C Superscript upper T Baseline plus upper R plus upper H upper Omega plus upper Omega Superscript upper T Baseline upper C Superscript upper T Baseline period EndLayout

With the previous modifications, the general UFIR filtering Algorithm 13 can be applied straightforwardly. To apply the GKF algorithm (3.160)–(3.165), noise vectors w Subscript k and v overbar Subscript k must be de‐correlated. Alternatively, to apply the GKF algorithm (3.146)–(3.151), a new optimal Kalman gain must be derived for time‐correlated noise.

11.5.2 Localization Performance

Tuning localization algorithms is an important process, which does not always end successfully due to the practical inability of measuring noise statistics in real time. To obtain reliable estimates over data arriving with a time step of tau equals 0.02 normal s, the best and worst cases were considered in [217].

Best Case: For the average quadrotor velocity of 0.6 normal m slash normal s estimated with a tolerance of 20%, the noise standard deviation is set as sigma Subscript w Baseline equals 0.12 normal m slash normal s. Noise in the first state is ignored, and the system noise covariance is specified with

upper Q equals Start 6 By 6 Matrix 1st Row 1st Column StartFraction tau squared Over 3 EndFraction 2nd Column StartFraction tau Over 2 EndFraction 3rd Column 0 4th Column 0 5th Column 0 6th Column 0 2nd Row 1st Column StartFraction tau Over 2 EndFraction 2nd Column 1 3rd Column 0 4th Column 0 5th Column 0 6th Column 0 3rd Row 1st Column 0 2nd Column 0 3rd Column StartFraction tau squared Over 3 EndFraction 4th Column StartFraction tau Over 2 EndFraction 5th Column 0 6th Column 0 4th Row 1st Column 0 2nd Column 0 3rd Column StartFraction tau Over 2 EndFraction 4th Column 1 5th Column 0 6th Column 0 5th Row 1st Column 0 2nd Column 0 3rd Column 0 4th Column 0 5th Column StartFraction tau squared Over 3 EndFraction 6th Column StartFraction tau Over 2 EndFraction 6th Row 1st Column 0 2nd Column 0 3rd Column 0 4th Column 0 5th Column StartFraction tau Over 2 EndFraction 6th Column 1 EndMatrix sigma Subscript w Superscript 2 Baseline period

The average noise variances in UWB data are measured as sigma Subscript v normal x Superscript 2 Baseline equals 0.16 normal m squared, sigma Subscript v y Superscript 2 Baseline equals 0.16 normal m squared, and sigma Subscript v z Superscript 2 Baseline equals 0.86 normal m squared. Accordingly, the measurement noise covariance is set as upper R equals diag left-parenthesis sigma Subscript v normal x Superscript 2 Baseline sigma Subscript v normal y Superscript 2 Baseline sigma Subscript v normal z Superscript 2 Baseline right-parenthesis. Experimentally, minimum localization errors were obtained by the optimal coloredness factor upper Psi Subscript opt Baseline approximately-equals 0.5.

Worst Case: Since the actual quadrotor velocity can vary by several times and the variance of UWB noise can be several times larger than in individual measurements, the error factor alpha was introduced as StartFraction 1 Over alpha squared EndFraction upper Q and alpha squared upper R, and it was supposed that in the worst case alpha equals 3 covers most of practical situations.

Schematic illustration of RMSEs produced by the KF, cKF, UFIR filter, and cUFIR filter. For the UWB boundary, filtering errors should be as low as possible.

Figure 11.17 RMSEs produced by the KF, cKF, UFIR filter, and cUFIR filter. For the UWB boundary, filtering errors should be as low as possible.

In this experiment, a quadrotor travels along a planned path considered as a reference trajectory. Typical RMSEs produced by the KF, GKF developed for CMN (cKF), UFIR filter, and general UFIR filter developed for CMN (cUFIR) are sketched in Fig. 11.17 along with the UWB error boundary. For this boundary, filtering errors should be as low as possible. What this experiment de facto reveals is that the properly tuned cKF is most accurate in the best case, but less accurate in the worst case. The cKF is also most sensitive to errors in the noise covariances, and its RMSE crosses the UWB boundary when alpha greater-than 2.4. The latter means that the more accurate cKF is less robust than the original KF. Thereby, this experiment confirms the earlier made conclusion that an increase in the estimation accuracy is achieved using additional tuning factors at the expanse of robustness. It is also seen that both UFIR filters are robust against changes in alpha and are reasonably accurate since their RMSEs do not exceed the UWB boundary.

11.6 Processing of Biosignals

Advances in sensor technology in recent decades have extended digital technologies to electrical and mechanical signals generated by the human body and referred to as biosignals. Such signals can be measured with or without contact with the body and are useful in a wide range of medical applications, from diagnostics and remote monitoring to effective prosthesis control. The best‐known bioelectrical signals are: electrocardiogram (ECG), electromyogram (EMG), electroencephalogram, mechanomyogram, electrooculography, galvanic skin response, and magnetoencephalogram. Depending on the nature of the origin, biosignals can be quasiperiodic, like an ECG, or impulsive, like an EMG.

Since biosignals are generated by the body at low electrical levels, their measurements are accompanied by intense noise of little‐studied origin. Another important specificity is that biosignals, as a rule, do not have simple models. Therefore, methods for optimal processing of biosignals are usually based on assumptions that cannot be theoretically substantiated, as in artificial intelligence. To illustrate possible extensions of the state estimation methods to biosignals, next we will consider examples of ECG and EMG biosignals and solve specific problems using Kalman and UFIR estimators.

11.6.1 ECG Signal Denoising Using UFIR Smoothing

The ECG signals allow detecting a wide variety of heart diseases. Since heartbeats may vary slightly from each other, accurate measurements are required. This is particularly important when data are used to extract ECG features and make decisions about the heart state using special software. Various types of noises contaminate the ECG signal during its acquisition and transmission, such as high frequency noise (electromyogram noise, additive Gaussian noise, and power line interference) and low frequency noise (baseline wandering). Because noise can cause misinterpretation of the heart state, efficient denoising is required.

Schematic illustration of single ECG pulse measured in the presence of noise.

Figure 11.18 Single ECG pulse measured in the presence of noise (Data). Noise reduction is provided by an UFIR smoother of degree i element-of left-bracket 2 comma 4 right-bracket.

A typical measured heartbeat pulse (Data) is shown in Fig. 11.18, where one can recognize the central rapid excursions, commonly referred to as the QRS‐complex, slow left excursion (T‐wave), slow first right excursion (T‐wave), and slow second right excursion (U‐wave). This suggests that the ECG signal mostly changes slow over time and is highly oversampled, with the exception of the QRS region, where it is critically sampled. Thus, in the frequency domain, the ECG signal energy is concentrated in two bands: low frequencies associated with LP filtering and high frequencies with BP filtering. By using the ith degree q‐lag UFIR smoother, suboptimal denoising of the slow part can be provided on the large horizon upper N Subscript opt, while the critically sampled fast excursions in the QRS‐complex require the shortest possible horizon upper N Subscript min Baseline equals i plus 1 to avoid output bias. To enable universal noise cancellation, an adaptive iterative q‐lag UFIR smoothing algorithm has been designed in [108].

Adaptive Smoothing of ECG Data

To adaptively remove noise from ECG data using an UFIR smoother, a window left-bracket normal upper Q Subscript int Baseline comma normal upper S Subscript int Baseline right-bracket is assigned to cover the QRS‐complex as shown in Fig. 11.18. Since points Q and S are well detectable in the ECG pulse, the window boundaries are assigned as normal upper Q Subscript int Baseline equals normal upper Q minus upper N Subscript opt and normal upper S Subscript int Baseline equals normal upper S plus upper N Subscript opt. The horizon upper N Subscript opt is measured by removing the QRS‐complex. By determining normal upper Q Subscript int, normal upper S Subscript int, upper N Subscript min, and upper N Subscript opt, the q‐lag UFIR smoothing filter provides the following estimates in the selected ECG regions:

  • ModifyingAbove x With tilde Subscript n vertical-bar n plus q Baseline left-parenthesis upper N Subscript opt Baseline right-parenthesis in left-bracket 0 colon normal upper Q Subscript int Baseline minus 1 right-bracket
  • ModifyingAbove x With tilde Subscript n vertical-bar n plus q Baseline left-parenthesis upper N Subscript adt Baseline equals upper N Subscript opt Baseline ellipsis upper N Subscript min Baseline right-parenthesis in left-bracket normal upper Q Subscript int Baseline comma normal upper Q minus 1 right-bracket
  • ModifyingAbove x With tilde Subscript n vertical-bar n plus q Baseline left-parenthesis upper N Subscript min Baseline right-parenthesis in left-bracket normal upper Q comma normal upper S minus 1 right-bracket
  • ModifyingAbove x With tilde Subscript n vertical-bar n plus q Baseline left-parenthesis upper N Subscript adt Baseline equals upper N Subscript min Baseline ellipsis upper N Subscript opt Baseline right-parenthesis in left-bracket normal upper S comma normal upper S Subscript int Baseline minus 1 right-bracket
  • ModifyingAbove x With tilde Subscript n vertical-bar n plus q Baseline left-parenthesis upper N Subscript opt Baseline right-parenthesis in left-bracket normal upper Q Subscript int Baseline colon End right-bracket

Here, the adaptive horizon upper N Subscript adt Baseline equals upper N Subscript opt Baseline ellipsis upper N Subscript min gradually decreases from upper N Subscript opt to upper N Subscript min with each new step, and upper N Subscript adt Baseline equals upper N Subscript min Baseline ellipsis upper N Subscript opt gradually increases from upper N Subscript min to upper N Subscript opt with each new step.

The efficiency of the low degree, i element-of left-bracket 2 comma 4 right-bracket, UFIR smoothing algorithm can be estimated from Fig. 11.18, and more details about this solution can be found in [108]. It is worth noting that Kalman smoothing cannot be effectively applied to suppress noise in ECG data due to unknown noise and rapid jumps in the QRS‐complex. Note also that in view of the ECG signal quasiperiodicity the harmonic state‐space model can also be used to provide denoising with UFIR smoothing.

11.6.2 EMG Envelope Extraction Using a UFIR Filter

EMG signals are records of motor unit action potentials (MUAPs) that reflect responses to electrical activity occurring in motor units (MUs) in a muscle. These signals are electrical, chemical, and mechanical in nature and are complex. Because MUAPs are acquired from different regions of the MUs, skeletal muscle contains hundreds of different types of muscle fibers, and the resulting signal is composed of several MUAPs. Measurements of EMG signals are usually performed in the presence of sensor noise and artifacts originating from various sources, such as electrocardiography interference, spurious background spikes, motion artifact, and power line interference. The MUAP characteristics depend, among other factors, on the geometric position of the electrode needle inserted into the muscle. By generating a MUAP, morphologic changes in muscle fibers can be detected and analyzed. More information on EMG signals can be found in [120] and reference therein.

Schematic illustration of EMG signal: (a) measured EMG signal u(t) composed by MUAPs, (b) Hilbert transform u^(t) of u(t) and envelope U(t)=u2(t)+u^2(t), and (c) extracted U(t) and desired EMG signal envelope.

Figure 11.19 EMG signal: (a) measured EMG signal u left-parenthesis t right-parenthesis composed by MUAPs, (b) Hilbert transform ModifyingAbove u With caret left-parenthesis t right-parenthesis of u left-parenthesis t right-parenthesis and envelope upper U left-parenthesis t right-parenthesis equals StartRoot u squared left-parenthesis t right-parenthesis plus ModifyingAbove u With caret squared left-parenthesis t right-parenthesis EndRoot, and (c) extracted upper U left-parenthesis t right-parenthesis and desired EMG signal envelope.

A typical measured EMG signal u left-parenthesis t right-parenthesis, composed by low‐density MUAPs, is shown in Fig. 11.19a. An important function of the EMG signal is the envelope, which is used in robotic systems and prothesis control to achieve proper human‐robot interaction. The envelope upper U left-parenthesis t right-parenthesis equals StartRoot u squared left-parenthesis t right-parenthesis plus ModifyingAbove u With caret squared left-parenthesis t right-parenthesis EndRoot can be shaped by combining u left-parenthesis t right-parenthesis with the Hilbert transform ModifyingAbove u With caret left-parenthesis t right-parenthesis as shown in Fig. 11.19b. However, this may not always lead to a smooth shape due to multiple ripples. It should be noted that for the sake of stability of the proportional control of the artificial prothesis, it is desirable that the envelope has a Gaussian shape (Fig. 11.19c).

To keep the EMG envelope as close to Gaussian as possible, it was proposed in [120] to treat the ripples on the envelope (Fig. 11.19c) as CMN and use a general UFIR filter. Accordingly, the oversampled envelope upper U Subscript k was represented by the upper K‐state polynomial model [172] under the Gauss‐Markov CMN v Subscript k assumption as

(11.48)y Subscript k Baseline equals upper H x Subscript k Baseline plus v Subscript k Baseline comma

where x Subscript k Baseline element-of double-struck upper R Superscript upper K is the upper U Subscript k state vector and y Subscript k is the scalar observation of upper U Subscript k. For the polynomial approximation, entries of the system matrix upper F are obtained using the Taylor series expansion as

(11.50)upper F equals Start 5 By 5 Matrix 1st Row 1st Column 1 2nd Column tau 3rd Column StartFraction tau squared Over 2 EndFraction 4th Column ellipsis 5th Column StartFraction tau Superscript upper K minus 1 Baseline Over left-parenthesis upper K minus 1 right-parenthesis factorial EndFraction 2nd Row 1st Column 0 2nd Column 1 3rd Column tau 4th Column ellipsis 5th Column StartFraction tau Superscript upper K minus 2 Baseline Over left-parenthesis upper K minus 2 right-parenthesis factorial EndFraction 3rd Row 1st Column 0 2nd Column 0 3rd Column 1 4th Column ellipsis 5th Column StartFraction tau Superscript upper K minus 3 Baseline Over left-parenthesis upper K minus 3 right-parenthesis factorial EndFraction 4th Row 1st Column vertical-ellipsis 2nd Column vertical-ellipsis 3rd Column vertical-ellipsis 4th Column down-right-diagonal-ellipsis 5th Column vertical-ellipsis 5th Row 1st Column 0 2nd Column 0 3rd Column 0 4th Column ellipsis 5th Column 1 EndMatrix

and the number of states upper K is chosen so that the best shaping is obtained.

The observation y Subscript k of upper U Subscript k corresponds to the first state. Therefore, the observation matrix is assigned as upper H equals left-bracket 1 0 ellipsis 0 right-bracket element-of double-struck upper R Superscript 1 times upper K. Matrix upper B element-of double-struck upper R Superscript upper K times upper P projects the upper U Subscript k noise w Subscript k Baseline element-of double-struck upper R Superscript upper P to x Subscript k. The scalar coloredness factor 0 less-than psi less-than 1 is determined during the testing phase to ensure the best shaping, and we notice that, by psi equals 0, noise v Subscript k becomes white Gaussian. Noise xi Subscript k is zero mean and white Gaussian, xi Subscript k Baseline tilde script í’© left-parenthesis 0 comma sigma Subscript xi k Superscript 2 Baseline right-parenthesis, with the variance upper E left-brace xi Subscript k Superscript 2 Baseline right-brace equals upper R Subscript k Baseline equals sigma Subscript xi k Superscript 2.

Since the noise in upper U Subscript k is nonstandard, it is assumed that w Subscript k has zero mean with uncertain statistics and distribution. To run KF, w Subscript k is treated as zero mean and white Gaussian, w Subscript k Baseline tilde script í’© left-parenthesis 0 comma upper Q right-parenthesis, with the covariance upper E left-brace w Subscript k Baseline w Subscript n Superscript upper T Baseline right-brace equals upper Q delta Subscript k minus n, where delta Subscript k is the Kronecker symbol, and the property upper E left-brace w Subscript k Baseline xi Subscript n Baseline right-brace equals 0 for all k and n. It is assumed that the estimate ModifyingAbove x With caret Subscript k of x Subscript k under the ripples in upper U Subscript k (Fig. 11.19c) will approach the Gaussian form if we assume that ripples are due to the CMN.

EMG Signals with Low‐Density MUAP

As an example, we consider an EMG signal composed of low‐density MUAPs that require the Hilbert transform to smooth the envelope [120]. The model in (11.47)(11.49) is defined by two states, upper K equals 2, and matrices

upper F equals Start 2 By 2 Matrix 1st Row 1st Column 1 2nd Column tau 2nd Row 1st Column 0 2nd Column 1 EndMatrix comma upper B equals StartBinomialOrMatrix StartFraction tau squared Over 2 EndFraction Choose tau EndBinomialOrMatrix comma upper H equals Start 1 By 2 Matrix 1st Row 1st Column 1 2nd Column 0 EndMatrix period

The EMG signal is taken from the “Elbowing” database [120], which contains samples collected from 11 subjects with knee abnormality previously diagnosed by a professional and 11 normal knees. Measurements are carried out using goniometry equipment MWX8 Datalog Biometrics with the sampling frequency f equals 1 kHz, which corresponds to tau equals 1 0 Superscript negative 3 Baseline normal s. Part of the database is shown in Fig. 11.20a. At the first step, the envelope is shaped using the Hilbert transform to represent “Data” as shown in Fig. 11.20b and Fig. 11.20c. Since information about noise is unavailable, the KF and UFIR filter are arbitrary tuned to produce consistent estimates with minimal variations regarding the desired smooth envelope and negligible time‐delays.

Schematic illustration of EMG signal composed with low-density MUAP and envelope extracted using the Hilbert transform (Data) [120]: (a) EMG signal, (b) KF and UFIR filtering estimate, and (c) KF and UFIR filtering estimates modified for CMN.

Figure 11.20 EMG signal composed with low‐density MUAP and envelope extracted using the Hilbert transform (Data) [120]: (a) EMG signal, (b) KF and UFIR filtering estimate, and (c) KF and UFIR filtering estimates modified for CMN.

The optimal horizon for a UFIR filter is measured as upper N Subscript opr Baseline equals 74, which means a high oversampling rate. To tune KF, several similar datasets are analyzed, and it is concluded that the assumed noise has a standard deviation of about sigma Subscript xi Baseline equals 50 mu upper V. Since the average third state (acceleration) of the envelope is about 0.5 normal upper V slash normal s squared with a tolerance of about 20 percent-sign or even more, the standard deviation of the process noise is set as sigma Subscript w Baseline equals 0.1 normal upper V slash normal s squared, which makes the KF estimate consistent with the UFIR estimate. With this filter setup, the extracted envelopes are sketched in Fig. 11.20b, and it turns out that even if the filters improve the envelope without significant time delays, there are still a lots of ripples, which require further smoothing.

To get rid of the ripples in the envelope, it is next assumed that the measurement noise is colored and of Gauss‐Markov origin (11.49). Then the coloredness factor is set as psi equals 0.75 to provide the best smoothing without significant time delay, the optimal horizon is measured as upper N Subscript opt Baseline equals 140, and the GKF and UFIR filter modified for CMN are applied. What follows from the result shown in Fig. 11.20c is that the filters essentially suppress ripples and improve the envelope and that the latter finally better matches the desired Gaussian shape.

11.7 Summary

In this chapter, we looked at several practical uses of FIR state estimators, and one can deduce that the FIR approach can be extended to many other traditional and new challenging tasks. The aim was to demonstrate the difference between the FIR and IIR (KF) estimates obtained from real data. The following important points should be kept in mind.

Optimal estimators such as the OFIR filter and KF serve well when the model is process‐fit and the noise is Gaussian or near Gaussian. Otherwise, the UFIR state estimator may be more accurate. There are some boundary conditions that separate the scope of optimal and unbiased estimates in terms of accuracy and robustness. But this is a subtle matter that requires investigations in individual cases. Indeed, it is always possible to set some noise covariance, even heuristically and with no justification, to make the KF accurate, and there is always an optimal horizon upper N Subscript opt that makes the UFIR estimate suboptimal. Therefore, the question arises, which is practically easier: to determine the noise covariances or upper N Subscript opt? The answer depends on the specific problem, but it is obvious that it takes less efforts to determine the scalar upper N Subscript opt than the covariance matrix. Moreover, given the horizon length of upper N, even not optimal, the UFIR state estimator becomes blind and therefore robust, which is highly required in many applications, especially industrial ones.

We conclude this chapter by listing several modern and nontraditional signal processing problems that can be solved using FIR state estimation technologies, and note that this list can be greatly expanded.

11.8 Problems

  1. ML is pursued to estimate parameters such as mean and covariance of medical images for normality and abnormality employing a support vector machine. Since noise in images is non‐Gaussian, it seems that robust UFIR/median smoothing may be more efficient.
  2. In uncertain random environments, the FIR approach can help to solve the problem of estimating instantaneous speed and disturbance load torque in motors with higher robustness than standard KF‐based algorithms.
  3. To adapt the KF‐based state estimator to uncertain environments under disturbances, artificial neural networks (ANNs) are used in what is referred to as a neuro‐observer, which is an EKF‐based structure augmented by an ANN to capture the unmodeled dynamics. It looks like a robust UFIR‐based observer with adaptive horizon might be a more efficient solution.
  4. To estimate state‐of‐charge of lithium‐ion batteries, which is critical for electrical vehicles, adaptive upper H Subscript infinity filtering is used to eliminate bias errors. Instead, the upper H Subscript infinity FIR and especially UFIR filtering algorithms should be tested as more robust and stable.
  5. To avoid the construction of new power plants, transmission lines, etc., the distributed generation (DG) approach has received a big attention. For optimal localization of several DGs under power losses, KF‐based algorithms are used. Since noise in a DG environment is typically non‐Gaussian, a robust UFIR state estimator may be the best choice.
  6. To estimate the size and location of a brain tumor, a massively parallel finite difference method based on the graphics processing unit is used together with a genetic algorithm to solve the inverse problem. To provide effective noise reduction on finite intervals, FIR state estimators can be built in to achieve greater accuracy.
..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset