logo资料库

迭代学习控制的基本概念及应用.pdf

第1页 / 共19页
第2页 / 共19页
第3页 / 共19页
第4页 / 共19页
第5页 / 共19页
第6页 / 共19页
第7页 / 共19页
第8页 / 共19页
资料共19页,剩余部分请下载后查看
Iterative learning control (ILC) is based on the notion that the performance of a system that executes the same task multiple times can be improved by learning from previous executions (trials, iterations, passes). For instance, a basketball player shooting a free throw from a fixed position can improve his or her ability to score by practicing the shot repeatedly. During each shot, the basketball player observes the trajectory of the ball and consciously plans an alteration in the shooting motion for the next attempt. As the player continues to practice, the cor- rect motion is learned and becomes ingrained into the muscle memory so that the shooting accuracy is iteratively improved. The converged muscle motion profile is an open-loop control generated through repetition and learn- ing. This type of learned open-loop control strategy is the essence of ILC. We consider learning controllers for systems that per- form the same operation repeatedly and under the same operating conditions. For such systems, a nonlearning con- troller yields the same tracking error on each pass. Although error signals from previous iterations are information rich, they are unused by a nonlearning controller. The objective of ILC is to improve performance by incorporating error information into the control for subsequent iterations. In doing so, high performance can be achieved with low tran- sient tracking error despite large model uncertainty and repeating disturbances. ILC differs from other learning-type control strate- gies, such as adaptive control, neural networks, and repeti- tive control (RC). Adaptive control strategies modify the controller, which is a system, whereas ILC modifies the control input, which is a signal [1]. Additionally, adaptive controllers typically do not take advantage of the information contained in repetitive com- mand signals. Similarly, neural network learning involves the modification of controller parameters rather than a control signal; in this case, large networks of nonlinear neurons are modified. These large networks require exten- sive training data, and fast convergence may be difficult to DOUGLAS A. BRISTOW, MARINA THARAYIL, and ANDREW G. ALLEYNE © DIGITALVISION & ARTVILLE A LEARNING-BASED METHOD FOR HIGH-PERFORMANCE TRACKING CONTROL 96 IEEE CONTROL SYSTEMS MAGAZINE » JUNE 2006 1066-033X/06/$20.00©2006IEEE
guarantee [2], whereas ILC usually converges adequately in just a few iterations. ILC is perhaps most similar to RC [3] except that RC is intended for continuous operation, whereas ILC is intend- ed for discontinuous operation. For example, an ILC appli- cation might be to control a robot that performs a task, returns to its home position, and comes to a rest before repeating the task. On the other hand, an RC application might be to control a hard disk drive’s read/write head, in which each iteration is a full rotation of the disk, and the next iteration immediately follows the current iteration. The difference between RC and ILC is the setting of the ini- tial conditions for each trial [4]. In ILC, the initial condi- tions are set to the same value on each trial. In RC, the initial conditions are set to the final conditions of the previ- ous trial. The difference in initial-condition resetting leads to different analysis techniques and results [4]. Traditionally, the focus of ILC has been on improving the performance of systems that execute a single, repeated operation. This focus includes many practical industrial systems in manufacturing, robotics, and chemical process- ing, where mass production on an assembly line entails repetition. ILC has been successfully applied to industrial robots [5]–[9], computer numerical control (CNC) machine tools [10], wafer stage motion systems [11], injection-mold- ing machines [12], [13], aluminum extruders [14], cold rolling mills [15], induction motors [16], chain conveyor systems [17], camless engine valves [18], autonomous vehi- cles [19], antilock braking [20], rapid thermal processing [21], [22], and semibatch chemical reactors [23]. ILC has also found application to systems that do not have identical repetition. For instance, in [24] an underwa- ter robot uses similar motions in all of its tasks but with different task-dependent speeds. These motions are equal- ized by a time-scale transformation, and a single ILC is employed for all motions. ILC can also serve as a training mechanism for open-loop control. This technique is used in [25] for fast-indexed motion control of low-cost, highly nonlinear actuators. As part of an identification procedure, ILC is used in [26] to obtain the aerodynamic drag coeffi- cient for a projectile. Finally, [27] proposes the use of ILC to develop high-peak power microwave tubes. The basic ideas of ILC can be found in a U.S. patent [28] filed in 1967 as well as the 1978 journal publication [29] written in Japanese. However, these ideas lay dormant until a series of articles in 1984 [5], [30]–[32] sparked wide- spread interest. Since then, the number of publications on ILC has been growing rapidly, including two special issues [33], [34], several books [1], [35]–[37], and two surveys [38], [39], although these comparatively early surveys capture only a fraction of the results available today. As illustrated in “Iterative Learning Control Versus Good Feedback and Feedforward Design,” ILC has clear advantages for certain classes of problems but is not applicable to every control scenario. The goal of the pre- Iterative Learning Control Versus Good Feedback and Feedforward Design The goal of ILC is to generate a feedforward control that tracks a specific reference or rejects a repeating distur- bance. ILC has several advantages over a well-designed feedback and feedforward controller. Foremost is that a feed- back controller reacts to inputs and disturbances and, there- fore, always has a lag in transient tracking. Feedforward control can eliminate this lag, but only for known or measur- able signals, such as the reference, and typically not for dis- turbances. ILC is anticipatory and can compensate for exogenous signals, such as repeating disturbances, in advance by learning from previous iterations. ILC does not require that the exogenous signals (references or distur- bances) be known or measured, only that these signals repeat from iteration to iteration. While a feedback controller can accommodate variations or uncertainties in the system model, a feedforward controller performs well only to the extent that the system is accurately known. Friction, unmodeled nonlinear behavior, and distur- bances can limit the effectiveness of feedforward control. Because ILC generates its open-loop control through prac- tice (feedback in the iteration domain), this high-performance control is also highly robust to system uncertainties. Indeed, ILC is frequently designed assuming linear models and applied to systems with nonlinearities yielding low tracking errors, often on the order of the system resolution. ILC cannot provide perfect tracking in every situation, however. Most notably, noise and nonrepeating disturbances hinder ILC performance. As with feedback control, observers can be used to limit noise sensitivity, although only to the extent to which the plant is known. However, unlike feedback control, the iteration-to-iteration learning of ILC provides opportunities for advanced filtering and signal processing. For instance, zero-phase filtering [43], which is noncausal, allows for high-frequency attenuation without introducing lag. Note that these operations help to alleviate the sensitivity of ILC to noise and nonrepeating disturbances. To reject nonre- peating disturbances, a feedback controller used in combina- tion with the ILC is the best approach. sent article is to provide a tutorial that gives a complete picture of the benefits, limitations, and open problems of ILC. This presentation is intended to provide the practic- ing engineer with the ability to design and analyze a sim- ple ILC as well as provide the reader with sufficient background and understanding to enter the field. As such, the primary, but not exclusive, focus of this survey is on single-input, single-output (SISO) discrete-time linear sys- tems. ILC results for this class of systems are accessible without extensive mathematical definitions and deriva- tions. Additionally, ILC designs using discrete-time JUNE 2006 « IEEE CONTROL SYSTEMS MAGAZINE 97
FIGURE 1 A two-dimensional, first-order ILC system. At the end of each iteration, the error is filtered through L, added to the previous control, and filtered again through Q. This updated open-loop control is applied to the plant in the next iteration. Error 0 1 1 Control Signal 0 0 1 j j+1 Iteration N−1 Sample Time + + Q P L j + j+1 − Disturbance Reference Iteration linearizations of nonlinear systems often yield good results when applied to the nonlinear systems [10], [22], [40]–[43]. ITERATIVE LEARNING CONTROL OVERVIEW Linear Iterative Learning Control System Description Consider the discrete-time, linear time-invariant (LTI), SISO system yj(k) = P(q)uj(k) + d(k), (1) where k is the time index, j is the iteration index, q is the forward time-shift operator qx(k) ≡ x(k + 1), yj is the out- put, uj is the control input, and d is an exogenous signal that repeats each iteration. The plant P(q) is a proper ratio- nal function of q and has a delay, or equivalently relative degree, of m. We assume that P(q) is asymptotically stable. When P(q) is not asymptotically stable, it can be stabilized with a feedback controller, and the ILC can be applied to the closed-loop system. Next consider the N-sample sequence of inputs and outputs uj(k), k ∈ {0, 1, . . . , N − 1}, yj(k), k ∈ {m, m + 1, . . . , N + m − 1}, d(k), k ∈ {m, m + 1, . . . , N + m − 1}, and the desired system output yd(k), k ∈ {m, m + 1, . . . , N + m − 1}. The performance or error signal is defined by ej(k) = yd(k) − yj(k). In practice, the time duration N of the trial is always finite, although sometimes it is useful for analysis and design to consider an infinite time duration of 98 IEEE CONTROL SYSTEMS MAGAZINE » JUNE 2006 N Time Sample infinite the trials [44]. In this work we use N = ∞ to denote trials with an infi- nite time duration. The iteration dimension indexed by j is usually considered with j ∈ {0, 1, 2, . . .}. For simplicity in this article, unless stated otherwise, the plant delay is assumed to be m = 1. Discrete time is the natural domain for ILC because ILC explicit- ly requires the storage of past-itera- tion data, which is typically sampled. System (1) is sufficiently general to capture IIR [11] and FIR [45] plants P(q). Repeating distur- bances [44], repeated nonzero initial conditions [4], and systems aug- mented with feedback and feedfor- ward control [44] can be captured in d(k). For instance, to incorporate the effect of repeated nonzero initial conditions, consider the system xj(k + 1) = Axj(k) + Buj(k) yj(k) = Cxj(k), (2) (3) with xj(0) = x0 for all j. This state-space system is equiv- alent to yj(k) = C(qI − A)−1B uj(k) + CAkx0 . P(q) d(k) Here, the signal d(k) is the free response of the system to the initial condition x0. A widely used ILC learning algorithm [1], [35], [38] is uj+ 1(k) = Q(q)[uj(k) + L(q)ej(k + 1)], (4) where the LTI dynamics Q(q) and L(q) are defined as the Q-filter and learning function, respectively. The two- dimensional (2-D) ILC system with plant dynamics (1) and learning dynamics (4) is shown in Figure 1. General Iterative Learning Control Algorithms There are several possible variations to the learning algo- rithm (4). Some researchers consider algorithms with lin- ear time-varying (LTV) functions [42], [46], [47], nonlinear functions [37], and iteration-varying functions [1], [48]. Additionally, the order of the algorithm, that is, the num- ber N0 and ei, i ∈ { j− N0 + 1, . . . , j}, used to calculate uj+ 1 can be increased. Algorithms with N0 > 1 are referred to as high- er-order learning algorithms [26], [36], [37], [44], [49]–[51]. The current error ej+ 1 can also be used in calculating uj+ 1 iterations of ui of previous
to obtain a current-iteration learning algorithm [52]–[56]. As shown in [57] and elsewhere in this article (see “Cur- rent-Iteration Iterative Learning Control”), the current-iter- ation learning algorithm is equivalent to the algorithm (4) combined with a feedback controller on the plant. Outline The remainder of this article is divided into four major sections. These are “System Representations,” “Analy- sis,” “Design,” and “Implementation Example.” Time and frequency-domain representations of the ILC sys- tem are presented in the “System Representation” sec- tion. The “Analysis” section examines the four basic topics of greatest relevance to understanding ILC sys- tem behavior: 1) stability, 2) performance, 3) transient learning behavior, and 4) robustness. The “Design” sec- tion investigates four different design methods: 1) PD type, 2) plant inversion, 3) H∞, and 4) quadratically optimal. The goal is to give the reader an array of tools to use and the knowledge of when each is appropriate. Finally, the design and implementation of an iterative learning controller for microscale robotic deposition manufacturing is presented in the “Implementation Example” section. This manufacturing example gives a quantitative evaluation of ILC benefits. SYSTEM REPRESENTATIONS Analytical results for ILC systems are developed using two system representations. Before proceeding with the analy- sis, we introduce these representations. Time-Domain Analysis Using the Lifted-System Framework To construct the lifted-system representation, the rational LTI plant (1) is first expanded as an infinite power series by dividing its denominator into its numerator, yielding P(q) = p1q −1 + p2q −2 + p3q −3 + ··· , (5) where the coefficients pk are Markov parameters [58]. is the impulse response. Note The sequence p1, p2, . . . that p1 = 0 since m = 1 is assumed. For the state space description (2), (3), pk is given by pk = CAk−1B. Stack- ing the signals in vectors, the system dynamics in (1) can be written equivalently as the N × N-dimensional lifted system Current-Iteration Iterative Learning Control Current-iteration ILC is a method for incorporating feedback with ILC [52]–[56]. The current-iteration ILC algorithm is given by uj+1(k ) = Q(q)[uj (k ) + L(q)ej (k + 1)] + C(q)ej+1(k ) and shown in Figure A. This learning scheme derives its name from the addition of a learning component in the current iteration through the term C(q)ej+1(k). This algorithm, however, is identical to the algorithm (4) combined with a feedback controller in the parallel architecture. Equivalence can be found between these two forms by separating the current-iteration ILC signal into feed- forward and feedback components as Feedback control uj+1(k ) = wj+1(k ) + C(q)ej+1(k ), where wj+1(k) = Q(q)[uj (k) + L(q)ej (k + 1)]. Then, solving for the iteration-domain dynamic equation for w yields wj+1(k) = Q(q)[wj (k) + (L(q) + q −1C(q))ej (k + 1)]. Therefore, the feedforward portion of the current-itera- tion ILC is identical to the algorithm (4) with the learn- ing function L(q) + q−1C(q). The algorithm (4) with learning function L(q) + q−1C(q) combined with a feedback controller in the parallel architecture is equiv- alent to the complete current-iteration ILC. Memory L Memory ILC Q wj uj G Plant yj yd − ej C Feedback Controller FIGURE A Current-iteration ILC architecture, which uses a control signal consisting of both feedforward and feedback in its learning algorithm. This architecture is equivalent to the algorithm (4) combined with a feedback controller in the parallel architecture. JUNE 2006 « IEEE CONTROL SYSTEMS MAGAZINE 99
  =   yj(1) yj(2) ... yj(N) yj   + 0 p1 p1 p2 ... ... pN pN−1   P   d(1) d(2) ... d(N) , ··· ··· . . . ···     0 0 ... p1   uj(0) uj(1) ... uj(N − 1) uj (6) and d =       ej(1) ej(2) ... ej(N) ej   −     . yj(1) yj(2) ... yj(N) yj yd(1) yd(2) ... yd yd(N) The components of yj and d are shifted by one time step to accommodate the one-step delay in the plant, ensuring that the diagonal entries of P are nonzero. For a plant with m-step delay, the lifted system representation is =     yj(m) yj(m + 1) yj(m + N − 1) ej(m) ej(m + 1) ej(m + N − 1) ... ...   =     = ×   −   ··· ··· . . . ··· 0 0 ... pm d(m) d(m + 1) ... d(m + N − 1)   , pm pm+1 ...  pm+N−1  uj(0) uj(1) ... 0 pm ...   pm+N−2  +      . uj(N − 1) yd(m) yd(m + 1) ...  yd(m + N − 1)  yj(m) yj(m + 1) ... yj(m + N − 1) The lifted form (6) allows us to write the SISO time- and iteration-domain dynamic system (1) as a multiple-input, multiple-output (MIMO) iteration-domain dynamic sys- tem. The time-domain dynamics are contained in the struc- ture of P, and the time signals uj, yj, and d are contained in the vectors uj, yj, and d. 100 IEEE CONTROL SYSTEMS MAGAZINE » JUNE 2006 Likewise, the learning algorithm (4) can be written in lifted form. The Q-filter and learning function can be non- causal functions with the impulse responses Q(q) = ··· + q−2q2 + q−1q1 + q0 + q1q −1 + q2q −2 + ··· and L(q) = ··· + l−2q2 + l−1q1 + l0 + l1q −1 + l2q −2 + ··· , respectively. In lifted form, (4) becomes     uj+1(0) uj+1(1) ... uj+1 uj+1(N − 1)   q0 q1 ... qN−1   l0 l1 ... lN−1 +           uj uj(0) uj(1) ... uj(N − 1)   .   ej(1) ej(2) ... ej(N) ej   (7) q−1 q0 ... qN−2 ··· ··· . . . ··· Q q−(N−1) q−(N−2) ... q0 l−1 l0 ... lN−2 ··· ··· . . . ··· L l−(N−1) l−(N−2) ... l0 When Q(q) and L(q) are causal functions, it follows that q−1 = q−2 = ··· = 0 and l−1 = l−2 = ··· = 0, and thus the matrices Q and L are lower triangular. Further distinctions on system causality can be found in “Causal and Non- causal Learning.” The matrices P, Q, and L are Toeplitz [59], meaning that all of the entries along each diagonal are identical. While LTI systems are assumed here, the lifted framework can easily accommodate an LTV plant, Q-filter, or learning function [42], [60]. The construction for LTV systems is the same, although P, Q, and L are not Toeplitz in this case. k=0 x(k)z Frequency-Domain Analysis Using the z-Domain Representation X(z) =∞ The one-sided z-transformation of the signal {x(k)}∞ k=0 is −1, and the z-transformation of a system is obtained by replacing q with z. The frequency response of a z-domain system is given by replacing z with ei θ for θ ∈ [−π, π]. For a sampled-data system, θ = π maps to the Nyquist frequency. To apply the z-transformation to the ILC system (1), (4), we must have N = ∞ because the
z-transform requires that signals be defined over an infi- nite time horizon. Since all practical applications of ILC have finite trial durations, the z-domain representation is an approximation of the ILC system [44]. Therefore, for the z-domain analysis we assume N = ∞, and we discuss what can be inferred about the finite-duration ILC system [44], [56], [61]–[64]. The transformed representations of the system in (1) and learning algorithm in (4) are Yj(z) = P(z)Uj(z) + D(z) (8) and Uj+1(z) = Q(z)[Uj(z) + zL(z)Ej(z)], (9) respectively, where Ej(z) = Yd(z) − Yj(z). The z that multi- plies L(z) emphasizes the forward time shift used in the learn- ing. For an m time-step plant delay, zm is used instead of z. ANALYSIS Stability The ILC system (1), (4) is asymptotically stable (AS) if there exists ¯u ∈ R such that |uj(k)| ≤ ¯u for all k = {0, . . . , N − 1} and j = {0, 1, . . . ,}, and, for all k ∈ {0, . . . , N − 1}, j→∞ uj(k) exists. lim We define the converged control as u∞(k) = limj→∞ uj(k). Time-domain and frequency-domain conditions for AS of the ILC system are presented here and developed in [44]. Substituting ej = yd − yj and the system dynamics (6) into the learning algorithm (7) yields the closed-loop itera- tion domain dynamics uj+1 = Q(I − LP)uj + QL(yd − d). (10) Let ρ(A) = maxi |λi(A)| be the spectral radius of the matrix A, and λi(A) the ith eigenvalue of A. The following AS condition follows directly. Theorem 1 [44] The ILC system (1), (4) is AS if and only if ρ(Q(I − LP)) < 1. (11) When the Q-filter and learning function are causal, the matrix Q(I − LP) is lower triangular and Toeplitz with repeated eigenvalues In this case, (11) is equivalent to the scalar condition |q0(1 − l0p1)| < 1. (13) Causal and Noncausal Learning O ne advantage that ILC has over traditional feedback and feedforward control is the possibility for ILC to anticipate and preemptively respond to repeated disturbances. This ability depends on the causality of the learning algorithm. Definition: The learning algorithm (4) is causal if uj+1(k) depends only on uj (h) and ej (h) for h ≤ k. It is noncausal if uj+1(k) is also a function of uj (h) or ej (h) for some h > k. Unlike the usual notion of noncausality, a noncausal learning algorithm is implementable in practice because the entire time sequence of data is available from all previous iterations. Consider the noncausal learning algorithm uj+1(k) = uj (k) + kpej (k + 1) and the causal learning algo- rithm uj+1(k) = uj (k) + kpej (k). Recall that a disturbance d(k) enters the error as ej (k) = yd(k) − P(q)uj (k) − d(k). There- fore, the noncausal algorithm anticipates the disturbance d(k + 1) and preemptively compensates with the control uj+1(k). The causal algorithm does not anticipate since uj+1(k) compensates for the disturbance d(k) with the same time index k. Causality also has implications in feedback equivalence [63], [64], [111], [112], which means that the converged con- trol u∞ obtained in ILC could be obtained instead by a feed- back controller. The results in [63], [64], [111], [112] are for continuous-time ILC, but can be extended to discrete-time ILC. In a noise-free scenerio, [111] shows that there is feed- back equivalence for causal learning algorithms, and further- more that the equivalent feedback controller can be obtained directly from the learning algorithm. This result suggests that causal ILC algorithms have little value since the same control action can be provided by a feedback controller without the learning process. However, there are critical limitations to the equivalence that may justify the continued examination and use of causal ILC algorithms. For instance, the feedback control equivalency discussed in [111] is limited to a noise- free scenario. As the performance of the ILC increases, the equivalent feedback controller has increasing gain [111]. In a noisy environment, high-gain feedback can degrade perfor- mance and damage equipment. Moreover, this equivalent feedback controller may not be stable [63]. Therefore, causal ILC algorithms are still of significant practical value. When the learning algorithm is noncausal, the ILC does, in general, preemptively respond to repeating disturbances. Except for special cases, there is no equivalent feedback controller that can provide the same control action as the converged control of a noncausal ILC since feedback control reacts to errors. λ = q0(1 − l0p1). (12) JUNE 2006 « IEEE CONTROL SYSTEMS MAGAZINE 101
Using (8), (9), iteration domain dynamics for the z-domain representation are given by Uj+1(z) = Q(z) [1 − zL(z)P(z)] Uj(z) + zQ(z)L(z) [Yd(z) − D(z)] . (14) Many ILC algorithms are designed to converge to zero error, e∞(k) = 0 for all k, independent of the refer- ence or repeating disturbance. The following result gives necessary and sufficient conditions for conver- gence to zero error. A sufficient condition for stability of the transformed sys- tem can be obtained by requiring that Q(z)[1 − zL(z)P(z)] be a contraction mapping. For a given z-domain system T(z), we define T(z)∞ = supθ∈[−π,π] Theorem 2 [44] If |T(eiθ )|. Q(z)[1 − zL(z)P(z)]∞ < 1, then the ILC system (1), (4) with N = ∞ is AS. (15) When Q(z) and L(z) are causal functions, (15) also implies AS for the finite-duration ILC system [44], [52]. The stability condition (15) is only sufficient and, in gener- al, much more conservative than the necessary and suffi- cient condition (11) [4]. Additional stability results developed in [44] can also be obtained from 2-D systems theory [61], [65] of which ILC systems are a special case [66]–[68]. Performance The performance of an ILC system is based on the asymp- totic value of the error. If the system is AS, the asymptotic error is e∞(k) = lim j→∞ ej(k) = lim j→∞(yd(k) − P(q)uj(k) − d(k)) = yd(k) − P(q)u∞(k) − d(k). Performance is often judged by comparing the difference between the converged error e∞(k) and the initial error e0(k) for a given reference trajectory. This comparison is done either qualitatively [5], [12], [18], [22] or quantitatively with a metric such as the root mean square (RMS) of the error [43], [46], [69]. If the ILC system is AS, then the asymptotic error is e∞ = [I − P[I − Q(I − LP)] −1QL](yd − d) (16) for the lifted system and E∞(z) = 1 − Q(z) 1 − Q(z)[1 − zL(z)P(z)] [Yd(z) − D(z)] (17) for the z-domain system. These results are obtained by replacing the iteration index j with ∞ in (6), (7) and (8), (9) and solving for e∞ and E∞(z), respectively [1]. 102 IEEE CONTROL SYSTEMS MAGAZINE » JUNE 2006 Theorem 3 [57] Suppose P and L are not identically zero. Then, for the ILC system (1), (4), e∞(k) = 0 for all k and for all yd and d, if and only if the system is AS and Q(q) = 1. Proof See [57] for the time-domain case and, assuming N = ∞, [11] for the frequency-domain case. Many ILC algorithms set Q(q) = 1 and thus do not include Q-filtering. Theorem 3 substantiates that this approach is necessary for perfect tracking. Q-filtering, however, can improve transient learning behavior and robustness, as discussed later in this article. More insight into the role of the Q-filter in performance is offered in [70], where the Q-filter is assumed to be an ideal lowpass filter with unity magnitude for low frequen- cies [0, θc] and zero magnitude for high frequencies θ ∈ (θc, π]. Although not realizable, the ideal lowpass filter is useful here for illustrative purposes. From (17), E∞(ei θ ) for the ideal lowpass filter is equal to zero for θ ∈ [0, θc] and equal to Yd(eiθ ) − D(eiθ ) for θ ∈ (θc, π]. Thus, for fre- quencies at which the magnitude of the Q-filter is 1, perfect tracking is achieved; for frequencies at which the magni- tude is 0, the ILC is effectively turned off. Using this approach, the Q-filter can be employed to determine which frequencies are emphasized in the learning process. Emphasizing certain frequency bands is useful for control- ling the iteration domain transients associated with ILC, as discussed in the “Robustness” section. Transient Learning Behavior We begin our discussion of transient learning behavior with an example illustrating transient growth in ILC systems. and learning Example 1 Consider the ILC system (1), (4) with plant dynamics yj(k) = [q/(q − .9)2]uj(k) algorithm uj+1(k) = uj(k) + .5ej(k + 1). The leading Markov parame- ters of this system are p1 = 1, q0 = 1, and l0 = 0.5. Since Q(q) and L(q) are causal, all of the eigenvalues of the lifted system are given by (12) as λ = 0.5. Therefore, the ILC system is AS by Theorem 1. The converged error is identi- cally zero because the Q-filter is unity. The trial duration is set to N = 50, and the desired output is the smoothed-step function shown in Figure 2. The 2-norm of ej is plotted in Figure 3 for the first 180 trials. Over the first 12 iterations, the error increases by nine orders of magnitude. Example 1 shows the large transient growth that can occur in ILC systems. Transient growth is problematic in
practice because neither the rate nor the magnitude of the growth is closely related to stability conditions. Recall that the eigenvalues of the closed-loop iteration dynamics in Example 1 are at λ = 0.5, which is well within the unit disk. Furthermore, it is difficult to distinguish transient growth from instability in practice because the initial growth rate and magnitude are so large. Large transient growth is a fundamental topic in ILC and preventing it is an essential objective in ILC design. Insights into the cause of large transient growth in ILC systems are presented in [4], [7], [43], [71], and [72]. To avoid large learning transients, monotonic conver- gence is desirable. The system (1), (4) is monotonically con- vergent under a given norm • if e∞ − ej+1 ≤ γe∞ − ej, for j ∈ {1, 2, . . .}, where 0 ≤ γ < 1 is the convergence rate. We now develop conditions for monotonic convergence. By manipulating the system dynamics (6), (7) and the asymptotic-error result (16), we obtain e∞ − ej+1 = PQ(I − LP)P −1(e∞ − ej). (18) When P(q), Q(q), and L(q) are causal (that is, P, Q, and L are Toeplitz and lower triangular), the matrices P, Q, and L commute, and (18) reduces to (e∞ − ej+1) = Q(I − LP)(e∞ − ej). [E∞(z) − Ej+1(z)] = Q(z) [1 − zL(z)P(z)] For the z-domain system, the error dynamics can be simi- larly obtained as E∞(z) − Ej(z) (19) Let ¯σ (·) be the maximum singular value and let · 2 denote the Euclidean norm. Using (18) and (19) we obtain the following monotonic convergence conditions. . Theorem 4 [44] If the ILC system (1), (4) satisfies −1 PQ(I − LP)P < 1, (20) = ¯σ γ1 then e∞ − ej+12 < γ1e∞ − ej2 for all j ∈ {1, 2, . . .}. Theorem 5 [11] If the ILC system (1), (4) with N = ∞ satisfies = Q(z) [1 − zL(z)P(z)]∞ < 1, γ2 (21) d y 1 0 0 10 20 30 40 50 Time (k) FIGURE 2 Reference command for Example 1. This smoothed step is generated from a trapezoidal velocity profile to enforce a bound on the acceleration. Industrial motion controllers often use smoothed references. 1010 105 100 2 | | j e | | 10−5 10−10 0 20 40 60 80 100 120 140 160 180 Iteration (j) FIGURE 3 Error 2-norm for Example 1. Despite the use of a smoothed step, the tracking error grows rapidly over the first ten then E∞(z) − Ej+1(z) ∞ < γ2 E∞(z) − Ej(z) ∞ for all j ∈ {1, 2, . . .}. When Q(z) and L(z) are causal functions, (21) also implies e∞ − ej+12 < γ2e∞ − ej2 for j ∈ {1, 2, . . .} for the ILC system with a finite-duration N [44]. Note that the z-domain monotonic convergence condition (21) is identi- cal to the stability condition (15) given in Theorem 2. Thus, when Q(z) and L(z) are causal functions, the stability con- dition (15) provides stability and monotonic convergence independent of the iteration duration N. In contrast, the monotonic convergence condition of the lifted system (20) is a more stringent requirement than the stability condition (11), and both are specific to the iteration duration N under consideration. In some cases, the learning transient behavior of an ILC system may be more important than stability. Some researchers have argued that unstable ILC algorithms can be effective if their initial behavior quickly decreases the error [52], [56], [71]. These algorithms can then be said to satisfy a “practical stability” condition because the learn- ing can be stopped at a low error before the divergent learning transient behavior begins. JUNE 2006 « IEEE CONTROL SYSTEMS MAGAZINE 103
分享到:
收藏