Automatica 47 (2011) 1007–1014

Contents lists available at ScienceDirect

Automatica journal homepage: www.elsevier.com/locate/automatica

Brief paper

Stability overlay for adaptive control laws✩ Paulo Rosa a,∗ , Jeff S. Shamma b , Carlos Silvestre a , Michael Athans a,c a

Institute for Systems and Robotics - Instituto Superior Tecnico, Av. Rovisco Pais, 1, 1049-001 Lisboa, Portugal

b

Georgia Institute of Technology, School of Electrical and Computer Engineering, Atlanta, GA, United States

c

EECS (emeritus), M.I.T., United States

article

info

Article history: Available online 25 February 2011 Keywords: Robust adaptive control Time-varying systems Robust stability

abstract This paper proposes an architecture referred to as Stability Overlay (SO) for adaptive control of a class of nonlinear time-varying plants. The SO can be implemented in parallel with a wide range of ‘‘performancebased’’ adaptive control laws, i.e., adaptive control laws that seek to improve closed-loop performance, but may be susceptible to instability in the presence of unaccounted model uncertainty. In this architecture, the performance-based adaptive control law designates candidate controllers based on performance considerations, while the SO supervises this selection based upon online robust stability considerations. A particular selection of a performance-based adaptive control law is not specified. Rather, this selection can be from a wide range of adaptive control schemes. This paper provides stability proofs for the SO architecture and presents a simulation that illustrates the applicability of the proposed method. © 2011 Elsevier Ltd. All rights reserved.

1. Introduction Adaptive control laws are required in many practical applications, where a single (non-adaptive) controller is not able to achieve the stability and/or performance requirements. This happens since every physical system can only be known up to some bound on the accuracy and, in particular, when the uncertain parameters vary with time. However, many adaptive control laws can lead to unstable closed-loop systems when connected to a plant with even the slightest discrepancies from the family of admissible plant models. This issue was first described in Rohrs, Valavani, Athans, and Stein (1985), in the so-called Rohrs et al. counterexample. Very small disturbances can be responsible for destabilizing the closed-loop due to the unavoidable unmodeled high frequency dynamics, present in every physical system.

✩ This work was partially supported by Fundação para a Ciência e a Tecnologia (ISR/IST pluriannual funding) through the POS_Conhecimento Program that includes FEDER funds, by the PTDC/EEA-ACR/72853/2006 OBSERVFLY Project, by the NSF project #ECS-0501394, and by the AFOSR project #FA9550-09-10538. The work of P. Rosa was supported by a Ph.D. Student Scholarship, SFRH/BD/30470/2006, from the FCT. The material in this paper was partially presented at the 2009 American Control Conference, June 10–12, 2009, St. Louis, Missouri, USA, and the 48th IEEE Conference on Decision and Control, December 16–18, 2009, Shanghai, China. This paper was recommended for publication in revised form by Associate Editor Alessandro Astolfi under the direction of Editor Andrew R. Teel. ∗ Corresponding author. Tel.: +351 218418054; fax: +351 218418291. E-mail addresses: [email protected] (P. Rosa), [email protected] (J.S. Shamma), [email protected] (C. Silvestre), [email protected] (M. Athans).

0005-1098/$ – see front matter © 2011 Elsevier Ltd. All rights reserved. doi:10.1016/j.automatica.2011.01.068

This article proposes a solution to the stability problem common to many closed-loop linear and nonlinear, time-invariant and time-varying, systems with performance-based adaptive control laws. The strategy developed herein, referred to as Stability Overlay (SO), takes into account both stability objectives — often robust to a very wide class of disturbances and model uncertainty — and performance requirements—that, in general, assume a stronger knowledge about the plant to be controlled. The algorithms presented in the sequel are based upon Al-Shyoukh and Shamma (2007, 2009) and were first introduced in Rosa, Shamma, Silvestre, and Athans (2009a,b). They assess the ‘‘rewards’’ received by each controller after its most recent utilization, without any prior information on the bounds of the exogenous disturbances and sensors noise. A control law is then disqualified or not, based upon its rewards, in a similar way to what is done in Angeli and Mosca (2002), Baldi, Battistelli, Mosca, and Tesi (2010), Chang and Safonov (2008), Fu and Barmish (1986), Manuelli, Cheong, Mosca, and Safonov (2007), Manuelli, Mosca, Safonov, and Tesi (2008), Morse, Mayne, and Goodwin (1992), Safonov and Tsao (1995, 1997), Stefanovic and Safonov (2008) and Wang, Paul, Stefanovic, and Safonov (2005, 2007) and in the references therein. However, in our approach, we suggest that the SO should only be responsible for the stability of the plant, and thus another algorithm should run in parallel in order to satisfy the posed performance requirements. Although the SO can be used as an adaptive control method per se, such an approach can lead to low levels of performance, as explained in the sequel. Therefore, our methodology differs from the aforementioned ones, in the sense that the rewards of the controllers are not used to decide which controller leads to the highest closed-loop performance, but rather to guarantee

1008

P. Rosa et al. / Automatica 47 (2011) 1007–1014

that a control law which is not able to stabilize the plant is not persistently selected. Other solutions, such as the Lyapunov-based approach presented in Angeli and Mosca (2002), rely on the model of the plant and hence require stronger assumptions than the ones presented in the sequel. Indeed, to the best of our knowledge, this paper presents, for the first time, a theoretical proof that one can, at least in some cases, detect and correct instability in adaptive control schemes for time-varying plants, without prior assumptions other than feasibility. For the proposed SO methodology, it is not required to know the plant model to be controlled nor the disturbance properties— see Martenson (1985) for a model-free adaptive controller for linear time-invariant systems. Still, as shown in Fu and Barmish (1986), it is clear that the performance of the closed-loop can be severely affected if no knowledge is available about the plant. Nonetheless, the model-free characteristic of the present method ensures robustness to several types of model uncertainty. In a sense, if the actual plant is close to a plant model in the family used to design the adaptive control law, then the adaptation runs as usual, without (or with minor) intervention of the SO. If, however, the actual plant or the properties of the disturbances do not match the ones used during the design, the closed-loop system may become unstable. Therefore, instead of blindly continuing to use the adaptation law, we assess the norm of the inputs and outputs of the system, and eventually switch to a stabilizing controller. Contrary to non-adaptive controllers, the decision subsystems of adaptive control laws are typically highly dependent on the model of the disturbances and on model uncertainty. Thus, in general, the design assumptions of these decision subsystems are more prone to failure than the control subsystems by themselves. Control strategies such as the Robust Multiple-Model Adaptive Control (RMMAC) (Athans, Fekri, & Pascoal, 2005; Fekri, Athans, & Pascoal, 2006a,b; Rosa, Athans, Fekri, & Silvestre, 2007), the Multiple Model Adaptive Control with Mixing (Kuipers & Ioannou, 2010), and the Unfalsified Control (Baldi et al., 2010; Manuelli et al., 2008; Safonov & Tsao, 1995, 1997; Wang et al., 2007), among others (Angeli & Mosca, 2002; Wang et al., 2005), have successfully shown high levels of performance against complexand real-valued uncertainties in the plant, although they can lead the closed-loop to instability if the uncertain parameters of the plant vary with time. This article, thus, proposes the use of the SO to ensure input/output stability of these kinds of methodologies when applied to time-varying plants. We emphasize that adaptive control strategies should be able to handle time-varying plants, since these methods are usually applied to plants with drifting parameters, and thus guaranteeing stability of plants with timefrozen parameters is typically not enough. Therefore, the SO can be seen as a safety device that can be used with many adaptive algorithms, achieving high levels of performance while providing robust stability guarantees for several different types of modeling errors. We show that the applicability of the SO is very wide, in the sense that it can be used in parallel with several adaptive control laws, as long as a set of natural assumptions is satisfied. This paper is organized as follows. We start by introducing the notation and formally posing the problem in Section 2. In Section 3, some properties of LTI systems are derived. The main result for LTI plants is presented in Section 4. In Section 5, simulation results of the Rohrs et al. counterexample are shown. In Section 6, the SO is extended to nonlinear time-varying plants. Finally, in Section 7, some conclusions are provided. 2. Preliminaries and notation We define |x| as the euclidian norm of x ∈ Rn , and ‖A‖2 as the induced norm of the matrix A, i.e., ‖A‖2 = supx̸=0 |Ax|/|x|. We

Fig. 1. Feedback interconnection between the plant (1) and the controllers Ki , selected through signal α(t ).

further define, for any σ > 0, ‖z |σ[t1 ,t2 ] ‖ = supτ ∈[t1 , t2 ] e−σ (t2 −τ ) |z (τ )|, and ‖z |[t1 ,t2 ] ‖ = supτ ∈[t1 ,t2 ] |z (τ )|. Although in the sequel we consider a much more general family of plant models, at this point we assume that the plant can be modeled by x˙ = Ax + Bu + F ξ ,

x(0) = x0

y = Cx + Gθ

  z=

y u

Cx + Gθ u







 =

(1a) (1b)

u = Kα(t ) y = Kα(t )

(1c)

0 z

Kα(t ) ∈ So := {K1 , K2 , . . . , KNc }.

(1d) (1e)

The output variables z (·) and y(·) can include performance outputs such as the ones obtained by filtering the plant output and the control input with the weights Wy (s) and Wu (s), respectively—see Zhou and Doyle (1997). x0 ∈ Rn is a fixed (but unknown) initial condition, ξ (·) ∈ L∞ is an exogenous disturbance, θ (·) ∈ L∞ is the measurement noise and u(·) is the control input. So is the set of eligible controllers which are considered to be constant matrix gains, without loss of generality, as explained in the sequel. N is the number of elements in So , and Ki , for i ∈ {1, 2, . . . , N }, represents a controller. Define a finitely switching control input as ufs (t ) =



Kα(t ) (y(t )), K∗ (y(t )),

0 ≤ t < to ; t ≥ to .

(2)

Fig. 1 depicts the output feedback interconnection between the plant and the controllers Ki , selected through signal α(t ). 3. Properties of LTI closed-loop systems In this section, we derive some properties of LTI closed-loop systems that are going to be useful next. Consider the LTI plant described by (1). Without loss of generality, we assume that the N controllers in So are static output feedback controllers, i.e., each Ki , for i ∈ {1, . . . , N }, is a constant matrix. Notice that, as shown in Al-Shyoukh and Shamma (2007, 2009) if the controllers are dynamic LTI systems, then switching between dynamic controllers can be represented, through appropriate state augmentation, as switching between ‘‘static output feedback’’ controllers. We follow closely the steps in Al-Shyoukh and Shamma (2007, 2009) to show that, under mild assumptions, (1) has the following properties: Property 1. For any finitely switching input (2) and any to , 1T , σ > 0, ‖z |σ[0,to ] ‖ < ∞ ⇒ ‖z |σ[0,to +1T ] ‖ < ∞. Property 2 (Input/Output Stability). There exist a control law, K∗ , and positive constants σ and l∗ , such that for any 0 < γ < 1, there exists a 1T ∗ ≥ 0 that satisfies the following condition. For any finitely switching control input (2), ‖z |σ[0,to +1T ] ‖ ≤ γ ‖z |σ[0,to ] ‖ + l∗ , for all 1T ≥ ∆T ∗ and to ≥ 0.

P. Rosa et al. / Automatica 47 (2011) 1007–1014

1009

other (probably performance-based) adaptive control algorithm, which may take into account the model of the plant. Regarding the initializations of the SO Algorithm #1, we can summarize the calculations suggested, based upon Section 3: choose arbitrary σ > 0; choose arbitrary positive γ < 1; choose values for 1To , lo based upon Rosa et al. (2009a)—the value of 1To is related to the slowest pole of the closed-loop system, while lo is related to the maximum amplitude of the exogenous disturbances. It is important to stress that smaller values of 1To and lo can actually lead to better transients, since we have only derived in Rosa et al. (2009a) (very conservative) upper bounds for those parameters. Nevertheless, these calculations should point out reasonable magnitudes for the parameters, whenever both the plant and the controllers are linear and time-invariant. Moreover, the stability guarantees are not affected by the selection of the initial values for these parameters. Fig. 2. Stability overlay (SO) Algorithm #1, for LTI plants.

Property 1 ensures there are no controllers in the legal set (referred to as eligible controllers) that take the output of the plant to infinity in finite time. Property 2 states that there is at least one eligible controller that satisfies a desired stabilization condition. Parameter l∗ accounts for the exogenous disturbances and the initial conditions. Consider the following assumption: Assumption 1. The switched linear system (1) satisfies:

• R{λj (A + BKi C )} < 0, ∀j , for some i ∈ {1, . . . , N }. • The pair [A, C ] is observable. • The exogenous disturbances and measurement noise are bounded by some (possibly unknown) constants ξ0 and θ0 , respectively, i.e., |ξ (·)| ≤ ξ0 , |θ (·)| ≤ θ0 . Proposition 1 (Rosa et al., 2009a). Under Assumption 1, the linear system (1) has the Properties 1 and 2. The discounted norm ‖z |σ[0,to +1T ] ‖ can be interpreted as a state-norm estimator (cf. Garcia & Mancilla-Aguilar, 2002; Hespanha, Liberzon, Angeli, & Sontag, 2005). Therefore, our decisions of disqualifying or not a controller, according to this interpretation, will be based upon the estimate of the norm of the actual state of the closed-loop system. 4. Stabilty overlay for LTI plants The reward, r (n), after using controller Kp (·) during the time interval tn−1 ≤ t < tn is defined as r ( n) =

1, 0,



‖z |σ[0,tn ] ‖ ≤ γ ‖z |σ[0,tn−1 ] ‖ + l(n) otherwise,

(3)

where γ is a fixed scalar with 0 < γ < 1 and l(n) is going to be specified next. Fig. 2 describes the Stability Overlay (SO) Algorithm #1, for LTI plants. The notation S = S \ K (n) means ‘‘the exclusion of element K (n) from S’’. The initial set of eligible control laws is denoted So , while Ko is the first control law selected, 1T (n) is the time-interval while the control law K (n) is used, lo is the initial value of l(n) in (3), and linc and 1Tinc are the increments for l(n) and 1T (n), respectively, whenever all the control laws have failed in their most recent utilization. The reasoning for the sequence of increasing dwell-times 1T (n) — also used, for instance, in Fu and Barmish (1986) — is that small values of 1T (n) may incorrectly lead to the ‘‘disqualification’’ of stabilizing controllers. As further stressed in the sequel, the selection of the control law among the set of eligible ones, S, can be done according to any

Remark 1. The number of controllers, N, is not relevant in terms of stability, as long as the initial set of control laws, So , contains a stabilizing controller. However, larger values of N may require longer searching periods, before a stabilizing controller is found. This obviously may deteriorate substantially the performance of the closed-loop. Theorem 1. If Assumption 1 is satisfied, the Stability Overlay Algorithm #1, for LTI plants, results in ‖z |σ[0,t ] ‖ bounded. Proof. We first recall Claim 1 in Al-Shyoukh and Shamma (2007, 2009): Claim 1. The parameters 1T (n) and l(n) are uniformly bounded. Proof. The parameters 1T (n) and l(n) are increased whenever every control law, Ki , in its most recent utilization resulted in a zero reward, i.e., ‖z |σ[0,tn ] ‖ > γ ‖z |σ[0,tn−1 ] ‖ + l(n). However, by Property 2, there exists a 1T ∗ ≥ 0, a positive constant l∗ , and at least one control law which receives a positive reward, provided that 1T (n) ≥ 1T ∗ . This implies that the condition r = 0 in (3) cannot be satisfied infinitely often with 1T (n) and l(n) increasing without bound.  According to Claim 1, there is at least one control law that is going to be used infinitely many times. For this control law, r = 1 in (3). Thus, all other control laws are going to be used at most a finite number of times. According to Proposition 1, the output is going to remain bounded during that (bounded) time interval where r = 0. For some to , the rewards obtained for t ≥ to are positive. Since the penalized output is bounded at t = to , it will remain bounded for t > to .  It is important to stress that we do not describe how to choose the controller to be put in the loop. In fact, we allow any controller in the set S (see Fig. 2) to be selected. The choice of the controller is responsible for the performance of the closed-loop and should be taken care of by some adaptive control law that (probably) takes into account the model of the plant and the disturbance properties. One example of the applicability of the SO with a model reference adaptive control architecture is presented in the following section. However, as shown in the sequel, the applicability of the SO is much wider. Many types of adaptive laws are eligible to be integrated with the SO, such as the schemes based on the identification of the plant parameters (see, for instance, Ioannou & Sun, 1996; Kuipers & Ioannou, 2008, and references therein), or the estimator-based methodologies in Hespanha et al. (2001). Remark 2. It should be noticed that the choice of the parameters for the algorithm may be very sensitive in some cases, depending upon the plant dynamics and the disturbances intensity. In fact,

1010

P. Rosa et al. / Automatica 47 (2011) 1007–1014

if the norm of the output of the closed-loop system grows very fast whenever a destabilizing controller is picked, and if the time required to disqualify a controller is very large, one may not get ‘‘practical stability’’. This means that, although a stabilizing controller is eventually selected, the transients may not be reasonable from a practical point of view. This effect may be exacerbated if we randomly pick the controllers in S, instead of using a performance-based algorithm to select among the eligible control laws. The reason for increasing 1T can be explained in a very intuitive manner, that relates it to the classical performance/robustness tradeoffs. In case 1T is very small, we may find the right controller faster, but we may also disqualify stabilizing controllers just because they were not used long enough. Therefore, large values of 1T guarantee stability at the cost of large transients, while smaller values of 1T can lead to smaller transients (higher performance) but may disqualify stabilizing controllers. Remark 3. The SO can also be interpreted in light of the so-called safe adaptive control considered in Baldi et al. (2010), Chang and Safonov (2008), Manuelli et al. (2007), Manuelli et al. (2008), Stefanovic and Safonov (2008) and Wang et al. (2007), all of which are based on the application of the Morse–Mayne–Goodwin Morse et al. (1992) hysteresis switching algorithm with cost-detectable cost functions. In fact, what we call here disqualified controllers is called falsified controllers in Wang et al. (2007), while the eligible/legal controllers correspond to the candidate controllers. The fact that a controller in So is able to stabilize the plant (Property 2) is referred to as the feasibilty condition in Wang et al. (2007). Finally, if we use the number of times the pair (1T (n), l(n)) has been incremented, as a cost function, then it would be cost-detectable in the sense of Wang et al. (2007), since this cost function would tend to infinity if the closed-loop is not stable—see Claim 1. Remark 4. Due to the nature of the adaptive control laws to which the SO can be applied, higher levels of performance than the ones achieved by the best controller, in the set of eligible ones, should not be expected. 5. Rohrs et al. counterexample To illustrate the usefulness of the SO for LTI plants, we use the so-called Rohrs et al. counterexample—see Rohrs et al. (1985). We use the reference model adaptation law referred to as ContinuousTime Algorithm 1 and the same terminology as in Rohrs et al. (1985). Further details regarding this simulation can be found in Rosa et al. (2009b). The first step in order to apply the SO is to discretize and bound the gains ky and kr in Rohrs et al. (1985), so that we have a finite number of controllers. Each pair (ky , kr ) defines a controller. Therefore, if ky is divided into ny bins and kr is divided into nr bins, the set So will have nr ny controllers. For this simulation, we use bins of width 2 and the limits are ±50. Fig. 3 shows that the closedloop is now stable, although some transients are also experienced, as explained in the sequel. The time instants when a controller is disqualified are also represented. If a disqualified controller is selected by the adaptive algorithm, the values of the adaptive gains kr and ky are updated to the legal ones closest to those obtained by the adaptive law. The bursting phenomenon observed in Fig. 3 is due to the switching among the controllers and to the use (during a certain amount of time) of controllers that are not able to stabilize the plant. This can be avoided, for instance, by resorting to fictitious reference signals (c.f., Angeli & Mosca, 2002; Chang & Safonov, 2008; Manuelli et al., 2007; Safonov & Tsao, 1997; Stefanovic & Safonov, 2008; Wang et al., 2005, 2007), which can be used to disqualify control laws without inserting them in the feedback loop.

Fig. 3. Output y(t ) of the closed-loop system using the continuous-time Algorithm 1, with the stability overlay. The red dashed lines indicate the time instants when the currently scheduled controller fails. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Fig. 4. Uncertainty region, Ω , for the parameter ρ , split into N subsets.

6. Stability overlay for time-varying plants The SO developed in Section 4 relies heavily on time-invariance for controller disqualification. In the case of time-varying plants, the algorithm in Fig. 2 will not work. Thus, in this section, a modification of the SO to accommodate slow time-variations of parameter drifting is presented. Consider a time-varying plant described by x˙ = f (x, u, w, ρ), y = g (x, u, w, ρ), z = h(x, u, w, ρ),

 P (ρ(t )) :=

x(0) = x0 , (4)

where x0 ∈ Rn is a fixed (but unknown) initial condition, w(·) ∈ L∞ is a bounded (but unknown) exogenous disturbance, ρ(·) is a vector of (possibly time-varying) parameters, u(·) is the control input and t ∈ R+ 0 explicitly denotes the time-dependence of the plant dynamics. We also define x(·) as the state of the plant and y(·) and z (·) as the measurement and the performance outputs, respectively. As for the LTI case, the variable z (·) can include performance outputs. We assume that ρ can be time-dependent and that it cannot be measured or estimated with the desired accuracy. For instance, let us consider the case where the process model has one parametric uncertainty, ρ ∈ [ρmin , ρmax ]. Although several switching MMAC methodologies are available in the literature to solve this kind of problem, they all share the same principles: in terms of design, we divide the (large) set of parametric uncertainty, Ω , into N (small) subregions, Ωj , j = {1, . . . , N } — see Fig. 4 — and synthesize a nonadaptive controller for them; in terms of implementation, we try to identify which region the uncertain parameter, ρ , belongs to, and then use the controller designed for that region. We are going to posit the following assumptions throughout the remainder of this article. Assumption 2. There exist continuous strictly increasing functions σ1 : R+ → R+ and σ2 : R+ → R+ and constant σ > 0, such that, for any finitely switching input (2), any 1T > 0, any to ≥ 0, and any (possibly time-varying) ρ(·),

‖z |σ[0,to +1T ] ‖ ≤ σ1 (1T )‖z |σ[0,to ] ‖ + σ2 (1T ).

P. Rosa et al. / Automatica 47 (2011) 1007–1014

1011

Assumption 2 severely constraints the class of nonlinear systems to which the SO can be applied. However, the main goal here is not to solve the general adaptive control problem for nonlinear plants, but rather to highlight the properties of the plant that are being exploited. For instance, in Kuipers and Ioannou (2010), the authors assume linear rational plants of known degree, admitting multiplicative modeling error, but also rational and with known slowest pole. In our approach, we do not use (explicitly) any such structural assumptions. Consider that ρ(t ) ∈ Ω ⊂ Rnρ for all t ≥ 0. The set Ω is the uncertainty region of the vector of parameters, ρ , as illustrated in Fig. 4. Let Ωj denote a subset of Ω . Assumption 3. Let Ωj ⊂ Rnρ , j = 1, 2, . . . , N satisfy Ω ⊂ j Ωj . There exist strictly positive constants l∗ , 1T ∗ , ν and γ , with γ < 1, such that, for any finitely switching input (2) with K∗ = Kj , if for all t ≥ to , (i) ρ(t ) ∈ Ωj and (ii) |ρ( ˙ t )| ≤ ν , then



‖z |σ[0,to +1T ] ‖ ≤ γ ‖z |σ[0,to ] ‖ + l∗ , for all 1T ≥ 1T ∗ and for σ as in Assumption 2. The value of ν is dictated by each controller and is referred to as the allowable time-rate of variation of the vector of parameters, ρ(·). The following definition is important when taking into account the allowable time-rate of variation of the dynamics of a plant. Definition 1. Suppose that ρ(tA ) ∈ Ωj . We say that the plant dynamics drifted at time instant t = tA if there exists δ ∗ > 0 such that, for every 0 < δ ≤ δ ∗ , we have ρ(tA + δ) ̸∈ Ωj . Furthermore, t = tA is referred to as a drifting time instant. Assumption 4. There exists Tmin > 0 such that, if ρ(t ) ∈ Ωj , then there exist t1 and t2 such that: (a) |t2 − t1 | ≥ Tmin ; (b) t1 ≤ t ≤ t2 ; (c) ρ(τ ) ∈ Ωj for all τ ∈ [t1 , t2 ].

Fig. 5. Stability Overlay (SO) Algorithm #2, for time-varying plants, with known 1T ∗ and l∗ . The notation S = S \ K (n) means ‘‘the exclusion of element K (n) from S’’.

According to Assumption 2, whenever a controller receives a zero reward, we have ‖z |σ[0,(i+1)1T ∗ ] ‖ ≤ Γ ‖z |σ[0,i1T ∗ ] ‖ + L, where Γ = σ1 (1T ∗ ) and L = σ2 (1T ∗ ). On the other hand, and according to Assumption 3, whenever a controller receives a positive reward, we have ‖z |σ[0,(i+1)1T ∗ ] ‖ ≤ γ ‖z |σ[0,i1T ∗ ] ‖ + l∗ . Then, if for a given integer n∗ , we have Tmin > (2N + n∗ )1T ∗ , we conclude that ∗ ‖z |σ[0,to +Tmin ] ‖ ≤ a‖z |σ[0,to ] ‖+ b, where a ≤ Γ 2N γ n and b is a func∗ ∗ tion of L, l , Γ , γ , N and n , since in every Tmin interval, a correct controller must be used at least n∗ times. Notice that, if a < 1, then ∃ : ‖z |σ[0,to ] ‖ ≤ z0 ⇒ ‖z |σ[0,to +Tmin ] ‖ ≤ z0 . z0 ≥0



Thus, if γ n <

1

the output of the system is bounded. There-

Hence, if tA and tB are drifting time instants with tA ̸= tB , then |tA − tB | ≥ Tmin . The reward after using controller Kp (·) is defined as in (3).

fore, a sufficient condition for Tmin is Tmin > (2N + logγ (Γ −2N )) 1T ∗ .

6.1. SO for TV Plants with 1T ∗ and l∗ Known

6.2. SO for TV plants with 1T ∗ and l∗ unknown

The Stability Overlay Algorithm #2, depicted in Fig. 5, for timevarying plants, with known 1T ∗ and l∗ , is described next. Contrary to the SO Algorithm #1, in this case we need not increase 1T and l, since they are assumed known a priori. Therefore, if all the controllers have received non-positive rewards, then the dynamics of the plant have drifted. Hence, a previously disqualified controller is the one which is able to stabilize the plant, and therefore we set S = So .

In this subsection, we present an alternative algorithm, that does not presume the knowledge of 1T ∗ and l∗ . Define z (l, ϵ) = l +ϵ , for γ as in Assumption 2 and ϵ > 0. Notice that, if ‖z |σ[0,t ] ‖ 1−γ

Theorem 2. Under Assumptions 2–4, for sufficiently large Tmin , the Stability Overlay Algorithm #2 for time-varying plants with known 1T ∗ and l∗ results in ‖z |σ[0,t ] ‖ bounded. The proof for sufficiently large Tmin is similar to the LTI case. In the sequel, an upper bound for Tmin is derived. In the algorithm depicted in Fig. 5, we suppose that 1T ∗ and l∗ are known (these can actually be upper bounds for 1T ∗ and l∗ , respectively). The difference between the algorithms for timeinvariant and for time-varying plants is that in the latter we never increase 1T and l, since we know 1T ∗ and l∗ , and these are the values used for every controller. Suppose that Tmin > 2N 1T ∗ . Then, during any time-interval Ta such that Ta ≥ Tmin , at least one controller receives a positive reward, even if the controllers are selected in a non-sequential fashion. To see this, consider that all the rewards we get during that time-interval are zero rewards. Then, all the N controllers have failed in a row, and, when connected again to the loop, they all failed once more. Since the plant dynamics can only drift once during the time-interval 2N 1T ∗ (see Definition 1), this is a contradiction.

≥ z ∗ :=

l∗ 1−γ

Γ 2N

, and a controller that receives a positive reward (3)

is selected, then, for 1T ≥ 1T ∗ , ‖z |σ[0,t +1T ] ‖ ≤ ‖z |σ[0,t ] ‖. In other words, for all ϵ > 0, if a controller that only receives positive re∗ wards (3) is used long enough, then, ‖z |σ[0,t +1T ] ‖ ≤ z + ϵ . Let So denote the set of available controllers for the SO. The Stability Overlay Algorithm #3, for time-varying plants with unknown 1T ∗ and l∗ , is shown in Fig. 6. This algorithm has two differences when compared to the timeinvariant case: (1) the sets of eligible controllers, S and Q , are ‘‘reset’’ (S = So and Q = So ) whenever the discounted norm of the output is below a given threshold, z (l(n), ϵ); (2) a controller can only be disqualified if it fails twice in between ‘‘resets’’ and ‖z |σ[0,t ] ‖ ≥ z (l(n), ϵ). As explained in the sequel, these modifications guarantee the boundedness of 1T (n) and l(n), while keeping the input/output stability of the closed-loop. Similarly to what happened with the SO Algorithm #1, the initializations 1To and lo can be performed arbitrarily, without affecting the stability of the closed-loop. The reasoning behind the use of set Q is as follows. Suppose that a given controller is not able to stabilize the plant and, hence, is disqualified. Then, suppose that the plant dynamics change and that the formerly disqualified controller is now the only eligible controller able to stabilize the plant. Using the architecture without set Q , we could only re-use previously disqualified controllers whenever we increased 1T (n) and l(n). Therefore, the set Q

1012

P. Rosa et al. / Automatica 47 (2011) 1007–1014

infinitely many times using this law, then limn→∞ 1T (n) → ∞, which is a contradiction. The same conclusion applies to l(n). In case (ii), we assume that the norm of the output decreases below z (l(n), ϵ) from time to time. In the sequel, we are going to show that, if 1T (n) (and, consequently, l(n)) is bounded, then ‖z |σ[0,t ] ‖ is also bounded. If 1T (n) is bounded, then, for some t2 ≥ 0, the update law 1T (n + 1) = 1T (n) + 1Tinc is not used for any t ≥ t2 . Let t˜ ≥ t2 be defined such that ‖z |σ[0,t˜] ‖ > z (l(n), ϵ). Further let t and t be

successive times such that t < t, ‖z |σ[0,t ] ‖ ≤ z (l(n), ϵ), ‖z |σ[0,t ] ‖ ≤ z (l(n), ϵ), and ∀ : ‖z |σ[0,t ] ‖ > z (l(n), ϵ). t ∈]t ,t¯[

Again, according to Assumption 2, for t ≥ t2 , whenever a controller receives a zero reward, we have ‖z |σ[0,t +1T ] ‖ ≤ Γ ‖z |σ[0,t ] ‖ + L, where Γ = σ1 (1T ) and L = σ2 (1T ), and according to Assumption 3, whenever a controller receives a positive reward, we have ‖z |σ[0,t +1T ] ‖ ≤ γ ‖z |σ[0,t ] ‖ + l∗ . Since we are assuming t ≥ t2 , we conclude that we cannot receive non-positive rewards (3) more than 2N − 1 times in the interval [t , t¯] (otherwise 1T (n) would be increased, which would be a contradiction). Therefore, the norm of the output at time t < t < t¯ is bounded by ‖z |σ[0,t ] ‖ ≤ a‖z |σ[0,t ] ‖ + b = az (l(n), ϵ) + b,

Fig. 6. Stability overlay (SO) Algorithm #3, for time-varying plants with unknown 1T ∗ and l∗ .

accounts for time-variations of the plant dynamics, since 1T (n) and l(n) are only increased whenever both sets, S and Q , are empty. Remark 5. The parameters l(n) and 1T (n) can also be reinitialized, whenever the condition ‖z |σ[0,t ] ‖ ≤ z¯ (l(n), ϵ) is satisfied. On one hand, this prevents the loss in terms of performance that can arise from the use of large values of 1T (n) and l(n), which in turn can be caused by occasional adverse disturbances. On the other, longer periods may be required to select a stabilizing controller whenever the plant parameters drift, since the values of 1T (n) and l(n) might have to be increased every time the condition ‖z |σ[0,t ] ‖ ≤ z¯ (l(n), ϵ) is satisfied. The following theorem summarizes the main result of this section. Theorem 3. Under Assumptions 2–4, for sufficiently large Tmin , the Stability Overlay Algorithm #3 for time-varying plants results in ‖z |σ[0,t ] ‖ bounded. Proof. For the sake of simplicity, we divide the proof of Theorem 3 into several small steps. We start by showing that bounded 1T (n) and l(n) imply bounded ‖z |σ[0,t ] ‖, and finally we show that 1T (n) and l(n) are bounded. Claim 2. If limn→∞ 1T (n) < ∞ and limn→∞ l(n) < ∞, then limt →∞ sup ‖z |σ[0,t ] ‖ < ∞. Proof. In words, this means that if ‖z |σ[0,t ] ‖ is unbounded, then 1T (n) (and, consequently, l(n)) is also unbounded. For the sake of simplicity of the proof, we consider the following cases separately: (i) ∃ : ∀ , ‖z |σ[0,t ] ‖ > z (l(n), ϵ); (ii) ∀ : ∃ , ‖z |σ[0,t ] ‖ ≤ to

t ≥ to

t1

t ≥ t1

z (l(n), ϵ). We prove case (i) by contradiction. Indeed, consider that ‖z |σ[0,t ] ‖ satisfies (i) and is unbounded, but that 1T (n) (and, consequently, l(n)) is bounded. In that case, we have that: (a) none of the controllers is persistently receiving positive rewards (3), since the norm of the output is not bounded, and (b) sets S and Q are never being reset, since ‖z |σ[0,t ] ‖ > z (l(n), ϵ) for all t ≥ to . Hence, at some point, S ∪ Q = ∅. Thus, 1T (n) is increased according to 1T (n + 1) = 1T (n) + 1Tinc . If 1T (n) is increased

where a = Γ 2N −1 γ n˜ , for some n˜ ≥ 1, and b is a positive constant (see the proof of Theorem 2). Since a ≤ Γ 2N −1 , we conclude that ‖z |σ[0,t ] ‖ can be bounded by a constant which is independent of the choice of t and t¯. Therefore, ‖z |σ[0,t ] ‖ is uniformly bounded, which concludes the proof.  Claim 3. Suppose that tA is a drifting time instant and that ‖z |σ[0,tA ] ‖ ≤ θ . Then, for sufficiently large Tmin , 1T and l, ‖z |σ[0,tA +to ] ‖ ≤ z (l, ϵ), for some to ∈ [0, Tmin ]. Proof (By Contradiction). Suppose Claim 3 is not true. Then, : ‖z |σ[0, tA +to ] ‖ > z (l, ϵ), which means that the ‘‘reset’’ ∀ to ∈[0, Tmin ]

S = So and Q = So will never occur. Hence, the algorithm behaves as the SO for LTI plants—see Section 4. However, the SO for time-invariant plants guarantees that, for sufficiently large to , ‖z |σ[0, tA +to ] ‖ ≤ z (l, ϵ), which is a contradiction.  Claim 4. Suppose that ρ(tA ) ∈ Ωj , where tA is a drifting time instant, and that, at t = tA , the selected control law is Kj (·). Then, for sufficiently large Tmin , 1T and l, and for sufficiently small δ > 0, we have S = So at time instant t = tA + Tmin − δ . Proof. Claim 4 states that before every drifting time instant and assuming Tmin , 1T and l large enough, we have S = So , where So is the set of all available controllers for the adaptive law. Using Claim 3, for large Tmin , : ‖z |σ[0,tA +to ] ‖ ≤ z (l, ϵ), we ∃ to ∈[0,Tmin ]

conclude that, for t = tA + to , we have S = So and Q = So . Since, for 1T ≥ 1T ∗ , there is at least one controller that receives a positive reward (3), for every bounded (but unknown) exogenous disturbance, w(·) ∈ L∞ , the set Q is never empty for t ∈ [tA + to , tA + Tmin ]. Thus, S = So at time instant t = tA + Tmin − δ , for some δ > 0.  Claim 5. The parameters 1T (n) and l(n) are uniformly bounded. Proof. Using Claim 4, we conclude that, for sufficiently large Tmin , 1T and l, Q = So at the drifting time instants and hence all the controllers are allowed to be connected to the loop. Since we are assuming Tmin , 1T and l are large enough, at least one of the controllers is going to receive a positive reward (3), for any w(·) ∈ L∞ . Given that 1T and l can only be increased if all the controllers fail, these parameters are uniformly bounded.  Using Claims 2 and 5, we conclude that the output of the timevarying closed-loop system is bounded. 

P. Rosa et al. / Automatica 47 (2011) 1007–1014

6.3. Computation of an upper bound for Tmin Let tA , tB and tC be consecutive drifting time instants, satisfying Assumption 4, i.e., min{|tA − tB |, |tB − tC |} ≥ Tmin . Also, consider 1T ≥ 1T ∗ and l ≥ l∗ . According to Claim 3, ‖z |σ[0,tA ] ‖ ≤ z (l, ϵ), 0

for some tA0 ∈ [tA , tB ]. Hence, similarly to what was derived in ∗ Section 6.1, we have ‖z |σ[0,tB ] ‖ ≤ γ m Γ 2N z (l, ϵ) + ψ =: z 1 , ∗ where ψ is a continuous function of l , L, Γ , γ and m∗ , and where m∗ is the number of time-intervals 1T needed for a controller receiving positive rewards (3) to reduce the discounted norm of the output from z 1 to z (l, ϵ). Tmin is some time interval large enough so that the stability of the algorithm is guaranteed. Thus, Tmin ≥ tB0 − tB , where tB0 ∈ [tB , tC ] with ‖z |σ[0,tB ] ‖ ≤ z (l, ϵ). Therefore,



Tmin ≥ 2N + logγ



1

γ 2N



0

1T . As a remark, typically 1T > 1T ∗ ,

which means Tmin must also be larger, as expected. For a simulation illustrating the usefulness of the SO for time-varying plants, the reader is referred to Rosa et al. (2009b). 7. Conclusions This paper proposed a strategy, referred to as Stability Overlay (SO), that provides input/output stability guarantees for a wide set of adaptive control schemes, when applied to linear and a class of nonlinear time-varying plants. We take advantage of the on-line evaluation of the selected control law to disqualify those that do not comply with the stability requirements. Unlike other adaptive control strategies, we take into account both stability — often robust to a very wide class of disturbances and model uncertainty — and performance requirements—that, in general, assume a stronger knowledge of the plant. As a caveat, the choice of the parameters for the algorithm may be very sensitive. In those situations, although a stabilizing controller is eventually selected, the transients may not be reasonable from a practical point of view. The solutions available in the literature for this issue are, in general, based on the falsification of control laws without inserting them into the feedback loop (c.f., Angeli & Mosca, 2002; Chang & Safonov, 2008; Manuelli et al., 2007; Safonov & Tsao, 1997; Stefanovic & Safonov, 2008; Wang et al., 2005, 2007). However, these approaches require the estimation of a fictitious reference signal, which may be a difficult task in particular for nonlinear plants. Acknowledgement We wish to thank our colleagues A. Pascoal, P. Aguiar, V. Hassani and J. Vasconcelos for the many discussions on the field of robust adaptive control. We would also like to acknowledge the reviewers for their insightful comments and suggestions. References Al-Shyoukh, I., & Shamma, J. S. (2007). Switching supervisory control using calibrated forecasts. In 46th IEEE conference on decision and control. December (pp. 705–716). Al-Shyoukh, I., & Shamma, J. S. (2009). Switching supervisory control using calibrated forecasts. IEEE Transactions on Automatic Control, 54(4), 705–716. Angeli, D., & Mosca, E. (2002). Lyapunov-based switching supervisory control of nonlinear uncertain systems. IEEE Transactions on Automatic Control, 47, 500–505. Athans, M., Fekri, S., & Pascoal, A. (2005). Issues on robust adaptive feedback control. In Preprints of the 16th IFAC world congress (pp. 9–39). Baldi, S., Battistelli, G., Mosca, E., & Tesi, P. (2010). Multi-model unfalsified adaptive switching supervisory control. Automatica, 46(2), 249–259. Chang, M. W., & Safonov, M. G. (2008). Unfalsified adaptive control: the benefit of bandpass filters. In AIAA guidance, navigation and control conf. and exhibit. Honolulu, HI. August. Fekri, S., Athans, M., & Pascoal, A. (2006a). Issues, progress and new results in robust adaptive control. International Journal of Adaptive Control and Signal Processing, 20, 519–579.

1013

Fekri, S., Athans, M., & Pascoal, A. (2006b). Robust multiple model adaptive control (RMMAC): a case study. International Journal of Adaptive Control and Signal Processing, 21(1), 1–30. Fu, M., & Barmish, B. R. (1986). Adaptive stabilization of linear systems via switching control. IEEE Transactions on Automatic Control, AC-31(12), 1097–1103. Garcia, R. A., & Mancilla-Aguilar, J. L. (2002). State-norm estimator of switched nonlinear systems. In 2002 American control conference. Hespanha, J. P., Liberzon, D., Morse, A. S., Anderson, B., Brinsmead, T., & deBruyne, F. (2001). Multiple model adaptive control, part 2: switching. International Journal of Robust and Nonlinear Control, 11(5), 479–496. Hybrid systems in control [special issue]. Hespanha, J. P., Liberzon, D., Angeli, D., & Sontag, E. D. (2005). Nonlinear normobservability notions and stability of switched systems. IEEE Transactions on Automatic Control, 50(2), 154–168. Ioannou, P., & Sun, J. (1996). Robust adaptive control. Prentice Hall, Inc.. Kuipers, M., & Ioannou, P. (2008). Practical robust adaptive control: benchmark example. In Proceedings of the 2008 American control conference (pp. 5168–5173). Kuipers, M., & Ioannou, P. (2010). Multiple model adaptive control with mixing. IEEE Transactions on Automatic Control, 55(8), 1822–1836. Manuelli, C., Cheong, S. G., Mosca, E., & Safonov, M. G. (2007). Stability of unfalsified adaptive control with non scli controllers and related performance under different prior knowledge. In Proc. of the 2007 European control conference. Kos, Greece. July (pp. 702–708). Manuelli, G., Mosca, E., Safonov, M. G., & Tesi, P. (2008). Unfalsified virtual reference adaptive switching control of plants with persistent disturbances. In Proc. IFAC world congress. Seoul, Korea. July (pp. 8925–8930). Martenson, B. (1985). The order of any stabilizing regulator is sufficient a priori information for adaptive stabilization. Systems and Control Letters, 6(2), 87–91. Morse, A. S., Mayne, D. Q., & Goodwin, G. C. (1992). Applications of hysteresis switching in parameter adaptive control. IEEE Transactions on Automatic Control, 37(9), 1343–1354. Rohrs, C. E., Valavani, L., Athans, M., & Stein, G. (1985). Robustness of continuoustime adaptive control algorithms in the presence of unmodeled dynamics. IEEE Transactions on Automatic Control, 30(9), 881–889. Rosa, P., Athans, M., Fekri, S., & Silvestre, C. (2007). Evaluation of the RMMAC/XI method with time-varying parameters and disturbance statistics. In Proceedings of the mediterranean conference on automation and control. MED07. Athens, Greece. June. Rosa, P., Shamma, J. S., Silvestre, C. J., & Athans, M. (2009a). Stability overlay for adaptive control laws applied to linear time-invariant systems. In Proceedings of the 2009 American control conference. June (pp 1934–1939). Rosa, P., Shamma, J. S., Silvestre, C. J., & Athans, M. (2009b). Stability overlay for linear and nonlinear time-varying plants. In Proceedings of the 48th IEEE conference on decision and control. December (pp. 2435–2440). Safonov, M. G., & Tsao, T.-C. (1995). The unfalsified control concept: a direct path from experiment to controller. In B. A. Francis, & A. R. Tannenbaum (Eds.), Feedback control, nonlinear systems and complexity. Berlin: Springer-Verlag. Safonov, M. G., & Tsao, Tung-Ching (1997). The unfalsified control concept and learning. IEEE Transactions on Automatic Control, 42(6), 843–847. Stefanovic, M., & Safonov, M. G. (2008). Safe adaptive switching control: stability and convergence. IEEE Transactions on Automatic Control, 53(9), 2012–2021. Wang, D., Paul, A., Stefanovic, M., & Safonov, M. (2005). Cost-detectability and stability of adaptive control systems. In Proceeding of the 44th IEEE international conference on decision and control. December (pp. 3584–3589). Wang, R., Paul, A., Stefanovic, M., & Safonov, M. G. (2007). Cost-detectability and stability of adaptive control systems. International Journal of Robust and Nonlinear Control, 17(5–6), 549–561. Zhou, K., & Doyle, J. C. (1997). Essentials of robust control. Prentice Hall.

Paulo Rosa received his degree of Licenciatura in Electrical and Computer Engineering, in 2006, from Instituto Superior Tecnico in Lisbon, Portugal, where he is currently enrolled as a Ph.D. candidate. His research interests include robust adaptive control of time-varying systems, guidance and control of autonomous vehicles, and fault detection and isolation methods.

Jeff Shamma’s general research area is feedback control and systems theory. He received a BS in Mechanical Engineering from Georgia Tech in 1983 and a Ph.D. in Systems Science and Engineering from the Massachusetts Institute of Technology in 1988. He has held faculty positions at the University of Minnesota, Minneapolis; University of Texas, Austin; and University of California, Los Angeles; and visiting positions at Caltech and MIT. In 2007, he returned to Georgia Tech where he is a Professor of Electrical and Computer Engineering and Julian T. Hightower Chair in Systems & Control.

1014

P. Rosa et al. / Automatica 47 (2011) 1007–1014

Carlos Silvestre received the Licenciatura degree in Electrical Engineering from the Instituto Superior Tecnico (IST) of Lisbon, Portugal, in 1987 and the M.Sc. degree in Electrical Engineering and the Ph.D. degree in Control Science from the same school in 1991 and 2000, respectively. Since 2000, he is with the Department of Electrical Engineering of Instituto Superior Tecnico, where he is currently an Assistant Professor of Control and Robotics. Over the past years, he has conducted research on the subjects of navigation guidance and control of air and underwater robots. His research interests include linear and nonlinear control theory, coordinated control of multiple vehicles, gain scheduled control, integrated design of guidance and control systems, inertial navigation systems, and mission control and real time architectures for complex autonomous systems with applications to unmanned air and underwater vehicles.

Michael Athans received his BSEE in 1958, MSEE in 1959 and Ph.D. in 1961, all from the Univ. of California at Berkeley. He was a faculty member in the MIT EE&CS department until his early retirement in 1998. Since then he has been an invited Research Professor in the Systems and Robotics Lab., Instituto Superior Tecnico in Lisbon, Portugal. He has coauthored 3 books, over 300 papers and has received numerous awards from IEEE, AACS, and IFAC.

Stability overlay for adaptive control laws

Feb 25, 2011 - revised form by Associate Editor Alessandro Astolfi under the direction of Editor. Andrew R. Teel. ∗ ...... and Computer Engineering, in 2006, from Instituto Supe- ... trical Engineering and the Ph.D. degree in Control Science.

958KB Sizes 2 Downloads 239 Views

Recommend Documents

Direct adaptive control using an adaptive reference model
aNASA-Langley Research Center, Mail Stop 308, Hampton, Virginia, USA; ... Direct model reference adaptive control is considered when the plant-model ...

Secure overlay cloud storage with access control and ...
We design and implement FADE, a secure overlay cloud storage system that achieves ... a more fine-grained approach called policy based file assured deletion, ...

Direct Adaptive Control using Single Network Adaptive ...
in forward direction and hence can be implemented on-line. Adaptive critic based ... network instead of two required in a standard adaptive critic design.

Feasibility Checks and Control Laws for Reconfigurations of ...
on semidefinite programming to check for the feasibility of a ..... IEEE International Conference on Robotics and Automation, Anchor- age, Alaska, May 2010.

Feasibility Checks and Control Laws for ...
Abstract—A multi-spacecraft system consists of wirelessly- connected spacecraft that share resources in Earth orbit or deep space. One of the enabling technologies of such a fractionated space architecture is cluster flight, which provides the clus

Feasibility Checks and Control Laws for Reconfigurations of ...
and Engineering, 3:4:360–371, 2006. [2] S. Boyd and L. Vanderberghe. Convex Optimization. Cambridge. University Press, 2004. [3] O. Brown and P. Eremenko ...

Adaptive Filter Techniques for Optical Beam Jitter Control
or repetitive devices (engines, actuators, electric motors, etc) onboard the platform ..... The path length of beam was approximately 1 meter, therefore μm and ...

Smooth Multicast Congestion Control for Adaptive ...
for multimedia transmission over wired and wireless networks. ... high value of the time interval between two consecutive RTCP .... (3). However, this assumption is not true for the Internet. ..... publication in the Journal of High Speed Networks.

an adaptive parameter control strategy for aco - Semantic Scholar
Aug 16, 2006 - 1College of Computer Science and Engineering, South China University of Technology, Guangzhou, P. R. ... one of the best performing algorithms for NP-hard problems ..... program tools, systems and computing machines.

Adaptive Control for a Discrete-time First-order ...
Adaptive Control for a Discrete-time First-order Nonlinear System with ... of non-parametric part is characterized by a Lipschitz constant. L, and the nonlinearity of ...

Adaptive Output-Feedback Fuzzy Tracking Control for a ... - IEEE Xplore
Oct 10, 2011 - Adaptive Output-Feedback Fuzzy Tracking Control for a Class of Nonlinear Systems. Qi Zhou, Peng Shi, Senior Member, IEEE, Jinjun Lu, and ...

Adaptive Learning Control for Spacecraft Formation ...
utilized to develop a learning controller which accounts for the periodic ... Practical applications of spacecraft formation flying include surveillance, passive ... linear control techniques to the distributed spacecraft formation maintenance proble

Several Algorithms for Finite-Model Adaptive Control
Mathematics of Control, Signals, and Systems manuscript No. (will be inserted by the ...... Safonov MG, Cabral B (2001) Fitting controllers to data. Systems and ...

A Receding Horizon Control algorithm for adaptive management of ...
Apr 22, 2009 - eters are refined and the management horizon advances. The RHC .... energy transport model was used to drive the RHC algorithm qk | k.

Adaptive Rate Control for Streaming Stored Fine ...
are thus suitable for servers that stream a large number of simultaneous ... of our heuristic when run on top of TCP to when run on top of popular ...... t = mC t = tmax end. ∆k+2 = ∆(tmax end ) = Figure 3: Optimal state graph G. 0. 50. 100. 150.

A Receding Horizon Control algorithm for adaptive management of ...
Apr 22, 2009 - managing soil irrigation with reclaimed water. The largest current .... To explain RHC, we .... information x(k) and the first control u*(k) is then used to calculate the state x(k ю 1) ...... Technical Report NCSU-IE Technical Report

Adaptive Wisp Tree - a multiresolution control structure for simulating ...
for simulating dynamic clustering in hair motion. F. Bertails1 .... The use of adaptive level of detail is a well known way to improve .... We call segment the links.

Toward Quantified Risk-Adaptive Access Control for ...
Toward Quantified RAdAC for Multi-tenant Cloud Computing. 3 computing, it is ... In most common definitions, cloud computing is comprised of three services.

Adaptive Learning Control for Spacecraft Formation ...
Practical applications of spacecraft formation flying include surveillance, passive ... consider that the inertial coordinate system {X,Y,Z} is attached to the center of ...

Supermedia Transport for Teleoperations over Overlay Networks
significantly reduce latency compared with available transport services. Keywords. Teleoperation, Overlay networks, Forward error correction. 1 Introduction and ...

Knowledge Delivery Mechanism for Autonomic Overlay Network ...
Jun 19, 2009 - KBN broker, termed the Trigger Broker. The Trigger Broker receives incoming subscriptions from the policy server. (dynamically derived from its policy set) and stores these in a local subscription table. When management state (event) m