1. Introduction
Disturbance rejection is a fundamental problem in control theory and practice. There are narrow-band disturbances in a surprisingly large number of applications, such as the hard disk drive [
1,
2], suspension system [
3,
4], and power active filter [
5,
6]. In some applications, the frequencies of the disturbances are unknown and time-varying. Thus, an adaptive narrow-band disturbance rejection design, which can accommodate to the frequency change and achieve high performance, is urgently needed.
The adaptive narrow-band disturbance rejection has been extensively studied, and many control methods have been proposed. The phase-locked loop (PLL)-based adaptive feedforward cancellation (AFC) [
7,
8], the internal model principle (IMP)-based adaptive regulation [
3,
9], and the disturbance observer-based control (DOBC) [
1,
4,
5,
10] are the main techniques. In the PLL method, both the magnitude and frequency can be estimated in real time. However, the frequency responses of the plant at all the concerned frequencies need to be known in advance for parameter adaptation, which limits its application to some complicated plants whose frequency responses are not available. The IMP is also a powerful design approach to reject narrow-band disturbances. According to the IMP, the controller should include the disturbance model and have resonant poles determined by the disturbance frequencies. Many adaptive regulation methods based on the IMP have been proposed under the assumption that the plants can be precisely modeled. However, the effects of model uncertainties cannot be ignored for some plants and must be taken into account together with the narrow-band disturbances. Although many researchers have considered using the adaptive control theory to solve the unknown disturbances as well as the model uncertainties, the system may be destabilized in the presence of unmodeled dynamics [
11]. The DOBC is a promising alternative since both the narrow-band disturbances, and the model uncertainties can be rejected in a unified and simple manner. Thus, we focus on the DOBC for the adaptive narrow-band disturbance rejection problem in this paper.
Among the various types of disturbance observers, the resonant observer (RO) is a fundamental technique for narrow-band disturbance estimation [
5,
12], whereas it cannot deal with model uncertainties. Different from the RO design, which is performed in the state space, the design of the Q-filter based disturbance observer (QBDO) is conducted in the frequency domain; although, it is model-dependent. In most of the literature on the QBDO design, only the narrow-band disturbance can be estimated [
1,
4], and the core idea is to modify the frequency response of the traditional low-pass Q-filter based on the disturbance model. Jia et al. [
1] designed the Q-filter as an adaptive band-pass filter with the frequency estimation. Since the designs of disturbance rejection and frequency estimation algorithms are performed separately, this scheme is categorized as an indirect method. In contrast, the direct method involves adapting the Q-filter directly to minimize the effect of the narrow-band disturbance [
4]. In addition to rejecting the narrow-band disturbance, the capability of attenuating the low-frequency wide-band uncertainty is retained [
13,
14]. A period time-delay element, which can be updated in real time, is cascaded to a standard low-pass Q-filter [
13]. However, additional phase delays are introduced at the disturbance frequencies by the low-pass filter, which is undesired, especially for the high-frequency disturbances. To overcome this deficiency, a novel Q-filter was proposed [
14]. The frequency response of this Q-filter is unity at the frequency of the narrow-band disturbance, and it retains the characteristic of a low-pass filter at other frequencies. However, it cannot provide the estimations of states.
In recent years, the extended state observer (ESO), which is the core of the active disturbance rejection control (ADRC) [
15,
16], has become a popular approach for disturbance rejection and has been applied successfully in many fields. Unlike the QBDO design, which requires a nominal model of the plant, cascaded integrators are selected as the nominal model and a total disturbance (including both the internal uncertainties and the external disturbances) concept is used in this scheme. Therefore, the ESO can be applied to complicated plants whose models are difficult to obtain [
17]. Various enhanced forms have been proposed to address specific challenges in disturbance rejection. Considering that the standard ESO [
16] is limited in performance when high-frequency measurement noise is fed into it, cascade-parallel ESOs [
18,
19] are proposed to maintain the good noise suppression of the cascaded ESO and overcome its weakness of poor disturbance rejection. To handle the unknown input gains, adaptive ESOs are developed to estimate the input gains, unmeasured states, and total disturbance simultaneously [
20,
21]. To further promote the anti-disturbance property, different nonlinear mechanisms are introduced to the ESO design. To improve the convergence speed and estimation accuracy, the generalized super-twisting technique is introduced to construct the finite-time-convergent ESOs [
22,
23], which ensure that the estimation error can converge to zero in finite time. Although the convergence time becomes finite, it is closely related to the initial condition and grows unboundedly with the increase in the initial error. To address this problem, fixed-time-convergent ESOs are designed [
24,
25]. Additionally, with regards to the peaking phenomenon of the ESO, fractional-order ESOs [
26] and switching ESOs [
27] are developed to improve the transient performance. Many researchers [
28,
29,
30] have investigated the performance of the generalized extended state observer (GESO) on the narrow-band disturbance rejection, but it is assumed that the
mth-order derivative of the disturbance is zero, which is obviously not reasonable for the sinusoidal disturbance. Instead of increasing the ESO order, Zheng et al. [
31] cascaded a phase compensator, which can provide a phase lead around the frequency of the narrow-band disturbance, to a low-pass ESO to compensate for the undesired phase delay at the high frequency. In these methods [
28,
29,
30,
31], the ESO is still in the form of a low-pass filter. Therefore, a high observer bandwidth is required to guarantee the rejection performance of the high-frequency narrow-band disturbance, which can affect the robustness to unmodelled dynamics and measurement noises.
In view of the above-mentioned facts, we propose a systematic control design method for adaptive narrow-band disturbance rejection based on a resonant generalized extended state observer (RGESO) and a robust frequency estimation method. The main contributions are
(i) The frequency response of the RGESO is shaped so that both the low-frequency wide-band uncertainties and the high-frequency narrow-band disturbances can be simultaneously rejected. Therefore, the high observer bandwidth necessary for an ordinary ESO is avoided, which can greatly improve the system robustness.
(ii) The bandwidths for the wide-band uncertainty and the narrow-band disturbance in the RGESO are represented by various parameters whose physical meanings are explicit, which can make the parameter tuning process quite straightforward. Furthermore, both parameter optimization and stability robustness analysis methods are provided to facilitate its application.
(iii) A robust adaptive notch filter (ANF) is proposed to estimate the unknown and time-varying disturbance frequencies and update the parameters of the RGESO. A direct algorithm is designed to avoid the computationally complicated factorization. Additionally, a recursive version of the Levenberg–Marquardt (LM) method is used to enhance the convergence.
The remainder of the paper is organized as follows. The design process of the RGESO is presented in
Section 2. In
Section 3, the algorithm for direct frequency estimation is proposed.
Section 4 discusses parameter tuning and robustness analysis methods. Extensive simulation results on a voice coil motor are provided in
Section 5. Finally, concluding remarks are given in
Section 6.
2. Ordinary GESO Design
Before introducing the design process of the RGESO, a brief description of the GESO is necessary to clearly illustrate the differences and advantages of the proposed method. Consider a general single-input–single-output plant which is uncertain and nonlinear as follows:
where
y is the controlled output,
u is the control input,
d represents the external disturbance, and
denotes the nominal value of the input gain. In the framework of ADRC, all the internal uncertainties and external disturbances are lumped into a total signal
f. Here,
f can be further divided into a low-frequency wide-band uncertainty
and a high-frequency narrow-band disturbance
.
For the considered plant, it is assumed that the relative order and an approximate estimate of are known. They are prerequisites essential for the design of ADRC and RGESO-based control discussed in this paper. For a wide range of systems, these parameters can be derived either through dynamic modeling or by utilizing identification techniques based on experimental data. Additionally, it is assumed that the number of the narrow-band disturbances in is known, which determines the number of the resonant terms in the RGESO designed subsequently. Finally, to augment the high-order derivatives of f as states, it is assumed that f is differentiable and let represent its ith-order derivative.
To reject the disturbance with high-order dynamics, the original plant can be augmented with
m additional states, including the total disturbance and its high-order derivatives. Then, (
1) can be reformulated into a state space form as follows:
where
,
,
, and
.
To estimate the total disturbance, a GESO [
28], which is in the form of the conventional Luenberger observer, can be designed as
where
is the estimation of
,
is the output estimation error, and
is the observer gain vector. To facilitate tuning,
can be designed based on the parameterized bandwidth
[
16].
Combining (
2) and (
3), the dynamics of the estimation error can be obtained as
where
. Taking the Laplace transformation of (
4), the transfer functions from the total disturbance to the estimation errors can be obtained as
where
is a high-pass filter whose bandwidth is determined by
.
For the above GESO,
should be much higher than the largest frequency of the disturbance. When dealing with disturbances that include high-frequency narrow-band components, the frequency response of the GESO is band-pass for all frequencies that are less than
. However, the high
will degrade the robustness to the unmodeled dynamics and amplify the sensitivity to the measurement noise. In fact, such a high
is unnecessary since only the band-pass around the frequency of the narrow-band disturbance is needed just as the resonant controller [
6,
32]. Considering this, a novel RGESO is proposed in this paper to effectively reject the high-frequency narrow-band disturbance with a moderate
.
4. ANF for Disturbance Frequency Estimation
In this section, the ANF is designed to provide real-time estimation of the instantaneous frequency (IF) of the narrow-band disturbance.The definition of the IF for a real sinusoidal signal is given as
where
is the phase of the signal. The ideal frequency response of an ANF satisfies
There are many forms of finite or infinite impulse response approximations of (
22). The minimal parameter ANF with zeros distributed on the unit circle [
34] is the most popular one, which can be expressed as
where
is the unit delay operator,
N is the number of sinusoidal disturbances,
(
and
are the notch frequency and the sampling time, respectively), and
(
) is the design parameter that determines the notch width and depth. When
,
is a stable filter. As
decreases, the notch width widens, resulting in a faster convergence speed but lower estimation accuracy. Due to the constraint that the zeros are on the unit circle, the coefficients of
have a mirror symmetric form which can be written as
where
. When the ANF input is
, the output prediction
can be expressed as
where
and
Each component of
is a polynomial of
.
In the traditional ANF design,
is a directly adapted parameter vector, and the complicated factorization is subsequently required to obtain the frequency estimation. However, in this paper, the parameter vector
is adapted directly. The adaptation of
aims to minimize the following weighed quadratic prediction-error criterion:
where
(
) is the forgetting factor. A small value of
implies that only recent data are included in the criterion. This allows a fast-varying
to be tracked properly, which results in a low bias error in the estimate. The drawback is that the estimate will be more influenced by the noise, which gives a high variance error in the estimate. Therefore, the selection of
is a tradeoff between how fast the filter should track the changes in and the energy of the noise. From (
28), we can see that the identification priority is given to the frequency component with larger power. Since there are both narrow-band and wide-band signals in the total disturbance, a band-pass filter is added before the disturbance estimation enters the ANF to eliminate the effects of the wide-band components. The identification of
is a highly nonlinear least square problem. To develop the LM algorithm, the error gradient should be determined first. The error gradient with respect to the parameter vector is
Differentiating both sides of (
25) can obtain
This result implies that the gradient vector is effectively the regression
filtered by
. Such filtering can improve the convergence of the algorithm. To reduce the computation complexity, the following filtered variables:
are introduced to calculate
. The term
in (
29) is a Jacobian matrix:
After some mathematical manipulations, the recursive expression can be obtained as
for
and
. The first two terms can be easily derived as
Thus, the partial derivatives can be calculated efficiently based on the recursion. To improve the convergence speed, the posteriori error
is used for the error prediction and the gradient calculation in (
25) and (
30), since it is expected to yield a better estimation than
.
is calculated using the latest parameters
as
where
and
After obtaining the error gradient, the next step is to design the recursive LM (RLM) algorithm to estimate the frequencies. The RLM method can be summarized as
where
,
,
is the approximation of the Hessian matrix, and
is a identity matrix. When
, the diagonal elements of
will be dominant, and thus, the parameters will adapt in the steepest descent direction with a short step length. The direction is acceptable, but the convergence speed may be very low. When
, the algorithm just becomes the SGN method, which may not achieve the convergence. The LM method can make a trade-off between the acceptance of the direction and the convergence speed through adding the term
. To avoid the matrix inversion in (
38), the approximation method is used as follows:
where
is a
matrix (
is the dimension of the parameter vector) with the first column being
and the second column being a
zero vector except for its
element being one, which can be represented as
and
satisfies
should be adapted on-line such that the cost function value in (
28) decreases using the RLM method. The predicted reduction of (
28) at time
t is
where
. The actual reduction
is defined as
Then, the value of
can be adjusted by the following rules:
where
(
) and
(
) determine the time and magnitude of the adjustment, respectively.
At the start of the data processing, it is advisable to apply the algorithm with wider notches (i.e., a smaller value of
), thus increasing filter sensitivity to the presence of input sine waves. After convergence, it is recommended that a larger
is used, which improves the asymptotic performance. Since the estimates of the variables in the gradient are initially quite poor, they should therefore be assigned a lower weight in the criterion compared to later measurements. Therefore, transient processes are designed for
and
to improve the convergence speed at the beginning as
where
and
are the desired steady values, and
and
are the time constants of the processes.
Although the RLM is designed to improve the convergence performance, convergence cannot always be ensured since both the signal and the algorithm are stochastic in the presence of the measurement noise. Once the estimation converges to a false value well beyond our design consideration, the closed-loop performance will be degraded, and the stability may even be violated. Considering that the reliability is the most important aspect in practice, monitoring the estimation results based on the prior knowledge about the frequency range is necessary. Thus, the following strategy is proposed to further enhance the convergence performance:
where
and
are the upper and lower bounds for
, respectively.
Until now, the complete form of the proposed control scheme has been presented.
6. Numerical Simulations and Discussion
In this section, numerical simulations are performed on the voice coil motor (VCM) model [
42] to find whether the proposed method can achieve the high-precision position control under the multi-band disturbances. The nominal model of the VCM is
where
,
, and
. The external disturbance is added in the input channel as
where
and
are the magnitude and frequency of each sinusoidal signal, respectively, and
represents a white Gaussian noise. In the following simulations, let
,
rad/s, and set the standard derivation of
as 0.3544. According to the control objective, the parameters in (
52) are selected as
where
, and
is chosen to make the regulation time be less than 0.03 s. Additionally, the closed-loop system is required to have stability robustness against the variations of narrow-band frequencies within
. To make it clearer, the block diagram of the overall control structure for the VCM, including the plant, the controller, the observer, and the adding disturbances, is presented in
Figure 2. MATLAB R2023a is used for conducting the simulations, and the simulations are performed on a computer equipped with an Intel i7 processor and memory of 16 GB.
6.1. Performance Test for the ANF
In this subsection, comparisons between the proposed ANF and the SGN-based ANF are performed to demonstrate the advantages of the RLM algorithm in terms of convergence. The periodic signal includes three sinusoidal components with
and
rad/s. The parameters for the two algorithms are presented in
Table 1, where
represents the vector of true values. To ensure a fair comparison, the common parameters have been selected identically for both algorithms. A large
can make
have a large update initially to reflect the little confidence in
. Therefore, we let
. Here, we let
to make the simulation scenario more challenging. The values of the remaining parameters are determined based on the value ranges of each parameter, the parameter tuning criteria, and extensive simulation results. The number of data samples is 10,000, and the convergence is defined as when the final estimation errors of all three frequencies are less than 5%. Without the online supervision strategy, 100 simulations are conducted for different SNRs and the numbers of convergence for the two algorithms are shown in
Table 2. It is clear that the convergence cannot be achieved for all the scenarios, whereas the RLM-based ANF is more advantageous especially when the SNR is relatively high. Furthermore, it is found that the convergence is guaranteed for all the simulations by using the supervision strategy.
Next, the simulations of tracking time-varying frequencies under different SNRs are performed. The standard deviation (SD)
of the estimation bias for 100 simulations is shown in
Table 3. The SD with and without the bracket correspond to the frequency variation rate of 2
and 5
, respectively. It is obvious that the estimation precision degrades with decreasing the SNR and increasing the variation rate of the frequency. When there is no parameter supervision, some estimations may be false due to the noise. However, the estimations can track the frequencies accurately in all scenarios with the addition of the supervision strategy.
6.2. Parameter Tuning and Stability Robustness Analysis Results
Suppose that the required closed-loop bandwidth is 100 rad/s, and then,
and
can be selected as
and
, respectively. Since the velocity signal can be measured directly, the RGESO is designed based on the first-order ESO as follows:
The problem in (
52) is solved using the interior point method, and the optimized control parameters are presented in
Table 4. The closed-loop performance transfer functions
and
are shown in
Figure 3. It can be seen that the required tracking bandwidth of 100 rad/s is realized, and the disturbance rejection specifications in (
52) are all met. Due to the decoupling characteristic of the 2-DOF scheme,
is almost the same with the desired closed-loop dynamics in (
47), even though the error feedback gains are selected first without considering the dynamics of the RGESO.
Figure 4 and
Table 5 show the effect of
on the closed-loop performance. According to
Figure 4, the tracking performance for different values of
is almost the same when
is large enough. For the disturbance rejection performance, the frequency responses around the narrow-band frequencies are very close, and they are mainly determined by the parameters of the resonant term, while in other frequencies, a higher
can achieve better disturbance attenuation performance. The expense of a higher
is a higher crossover frequency and thus a smaller phase margin (as well as the time-delay tolerance), as can be seen in
Table 5. Since the optimized
can effectively reject the wide-band uncertainty and the narrow-band disturbance, a higher
is unnecessary and may degrade the robustness.
Next, the verification of the stability robustness within the variation range
of the frequency is performed based on the above-mentioned graphical method. The closed-loop characteristic polynomial can be obtained as
where
The value sets of the polynomial family are depicted in
Figure 5. It can be observed that the origin is not included in the value sets, and thus, the system is robustly stable.
6.3. Comparative Analysis
To illustrate the advantages of the proposed scheme, comparisons with
, whose order is
i , are performed. The error feedback gains are selected to be the same as those designed above for a fair comparison. For
, the transfer function from the practical disturbance to the disturbance estimation is
To perform a fair comparison between the
and the proposed RGESO, the bandwidths of the
are determined to achieve specified disturbance attenuation performance (as defined in (
52)) as
where
is the transfer function from the disturbance to the output. For the closed-loop systems designed based on the
,
can be obtained as
Because
is in the form of a high-pass filter, only the following condition for the highest narrow-band disturbance is needed to be satisfied:
To reduce
as much as possible, we let
Based on (
66), we can obtain
for
by solving the high-order equation in terms of
. After some calculations, the required bandwidths for
are derived and presented in
Table 4. The required bandwidth of
is the smallest, which is consistent with the previous results [
28,
29,
30]. The frequency responses of closed-loop transfer functions from the disturbance to the output are presented in
Figure 6. We can see that the disturbance attenuation requirement defined in (
52) is satisfied for the proposed and the comparative methods.
For RGESO- and
-based control systems, the transfer functions from the reference signal to the output are depicted in
Figure 7. Due to the sufficiently high bandwidths of RESO and
, the decoupling characteristic of the 2-DOF scheme is achieved, resulting in
of all methods closely resembling the desired closed-loop dynamics.
To evaluate the sensitivity to the measurement noise,
(the transfer function from the measurement noise to the control input) is analyzed. The frequency-domain responses of
for various methods are shown in
Figure 8. As evident in
Figure 8, the significantly higher bandwidths of
result in a substantial amplification of the noise in the control signal, potentially accelerating actuator wear and tear. By contrast, the RESO causes the least amount of amplification on noise, indicating its superior robustness to measurement noise.
To achieve the desirable attenuation magnitude at the narrow-band frequencies, the bandwidths have to be designed to be sufficiently large. With such high bandwidths, the robustness to the unmodeled dynamics (such as the time-delay) will be degraded. In addition, a high sampling rate is required, which is expensive and often unimplementable in practice. The gain margin (
), phase margin (
), crossover frequency (
), and time delay tolerance
(
), as well as the required Shannon sampling frequency
are shown
Table 6. It can be seen that increasing the order of GESO can help enhance the robustness to time delay and lower the required sampling time. However, the performance of the
is still poor compared with the RGESO.
To validate the robustness of various methods, we consider parameter perturbations and time delay. First, we performed 1000 Monte Carlo simulations, with the parameters (, , and ) randomly perturbed between −50% and 50% in each simulation, including the cases of extreme perturbations. The simulation results indicate that all methods are capable of maintaining closed-loop stability. Next, to simulate the effects of time delay, the VCM model is modeled as , where represents the time delay. Our analysis revealed that the closed-loop systems become unstable when . Notably, the proposed method exhibits the strongest robustness to a time delay.
6.4. Numerical Simulations
Numerical simulations are performed with the optimized controller. The sampling time is 0.0001 s. The disturbance frequencies are set as nominal values at the beginning and then have step changes to the maximal and minimal values at
s and
s, respectively. Only the PD (proportional derivative) control is employed in the first three seconds, and the disturbance estimation is added to the control signal thereafter. The simulation results are shown in
Figure 9,
Figure 10 and
Figure 11. As can be seen in
Figure 9, the disturbance rejection performance is greatly improved after using the disturbance estimation in the control signal calculation. Furthermore,
Figure 9 depicts the outputs with and without the adaptation to the frequency variation, and it can be found that the adaptation is necessary to achieve better disturbance attenuation. The tracking error increases slightly with the increase in the disturbance frequency since the attenuation magnitude becomes smaller at a higher frequency. The disturbance and frequency estimations are presented in
Figure 10 and
Figure 11, respectively. According to
Figure 10, the disturbance estimation can track the true value accurately with very small phase delay. In addition, the high-frequency noise can be filtered out effectively due to a moderate selection of
. As shown in
Figure 11, the frequency estimations can track the true values fast and accurately, which provides an important precondition for effective rejection of time-varying narrow-band disturbances.