Automatica 36 (2000) 5}22 Stable adaptive control with recurrent networksq Grzegorz J. Kulawski!,*,1, Mietek A. BrdysH " !Shell International Exploration and Production B.V., Research and Technical Services, Volmerlaan 8, P.O. Box 60, 2280 AB Rijswijk, The Netherlands "School of Electronic and Electrical Engineering, The University of Birmingham, Edgbaston, Birmingham B15 2TT, UK Received 4 November 1997; revised 9 December 1998; received in "nal form 6 March 1999 ¹he paper describes an adaptive control scheme for uncertain nonlinear plants with unmeasurable state, based on dynamic neural networks. ¹heoretical stability analysis and simulation examples are presented. Abstract An adaptive control technique for nonlinear stable plants with unmeasurable state is presented. It is based on a recurrent neural network employed as a dynamical model of the plant. Using this dynamical model, a feedback linearizing control is computed and applied to the plant. Parameters of the model are updated on-line to allow for partially unknown and time-varying plant. The stability of the scheme is shown theoretically, and its performance and limitations of the assumptions are illustrated in simulations. It is argued that appropriately structured recurrent neural networks can provide conveniently parameterized dynamic models for many nonlinear systems for use in adaptive control. ( 1999 Elsevier Science Ltd. All rights reserved. Keywords: Dynamic neural networks; Nonlinear adaptive control; Learning 1. Introduction From the point of view of algorithm design, universal approximation properties of static networks, shown for example in Funahashi (1989), Cybenko (1989) and Hornik, Stinchcombe and White (1989), make them a potentially useful tool for nonlinear control and identi"cation problems. In addition to that, their massively parallel structure combined with simplicity of single neurons brings bene"ts related to implementation issues such as speed of computations and fault tolerance due to natural redundancy. Realisation of these facts sparked recently a great deal of research activity in the control community for which the early paper (Narendra & Parthasarathy, 1990) takes much of the credit. A number of q This paper was not presented at any IFAC meeting. This paper was recommended for publication in revised form by Associate Editor A.E. de Barros Ruano under the direction of Editor C.C. Hang. * Corresponding author. Tel.: 31-70-311-6089; fax: #31-70-3112521. E-mail addresses: [email protected] (G.J. Kulawski), [email protected] (M.A. BrdysH ) 1 The paper was written during the author's stay at The University of Birmingham. di!erent control structures utilising neural networks have been subsequently proposed, while multilayer perceptron (MLP) and radial basis function (RBF) networks became the most popular neural architectures used. After the early enthusiasm and wealth of ideas, the challenge was to put neural networks more "rmly into the control engineering context and rigourously address such essential issues as guarantees of performance, robustness or stability. Control algorithms, in which neural networks are trained o!-line prior to their application in the on-line operation, when no further weights adjustments are made, are relatively simpler from the viewpoint of mathematical analysis. This approach was for example taken in two papers (Levin & Narendra, 1993, 1996), where a number of control methods using multilayer perceptron trained o!-line are presented and supported by solid mathematical treatment. Design of provably stable adaptive neural controllers posed more di$cult theoretical questions. One of the early adaptive schemes with local stability results has been presented in Chen and Khalil (1991). A globally stable nonlinear adaptive controller using RBF networks was proposed in Sanner and Slotine (1992). The fact that an output of a RBF network is linear in the adjustable parameters proved important for the design of a stable parameter update law. Finally, global stability has also 0005-1098/00/$ - see front matter ( 1999 Elsevier Science Ltd. All rights reserved. PII: S 0 0 0 5 - 1 0 9 8 ( 9 9 ) 0 0 0 9 2 - 8 6 G.J. Kulawski, M.A. Brdys& / Automatica 36 (2000) 5}22 been shown for an adaptive controller using a single hidden layer MLP network, in which both output and hidden weights are tuned, "rst in Lewis, Yesildirek and Liu (1993) and later further developed in Chang, Fu and Yang (1996). These developments are very much based on results from the theory of adaptive control of linear plants (see, for example Narendra & Annaswamy, 1989). In the case of linearly parameterised RBF networks the existing adaptive schemes could be applied in a relatively straightforward manner, while the nonlinear parameterisation of MLP networks was a signi"cant obstacle. It was then realised that, due to the speci"c properties of sigmoid activation functions, control laws, with approximations of unknown functions realised by MLP networks, can be "t into a framework which allows the use of methods from adaptive control for linear systems. While signi"cant progress has been achieved in applications of static neural networks for control, these algorithms are limited to the case when full state of the controlled plant is available for measurement. The output feedback control (when the full state is not measurable) and especially adaptive output feedback control for nonlinear systems remains one of the outstanding problems which are, at present, very actively researched. Systematic design techniques have only been developed for speci"c classes of nonlinear systems, e.g. Marino and Tomei (1995) and KrsticH , Kanellakopoulos and KokotovicH (1995). Lack of state measurements makes the problem particularly di$cult, as some kind of dynamic model of the system needs to be used for control law synthesis. For linear systems with known parameters, owing to the separation principle, the output feedback control design can be conveniently split into an independent design of a stable Luenberger observer and a stable state feedback law. However, there is no similar separation principle in the adaptive context even for linear systems, let alone nonlinear ones. In case of adaptive control of linear plants, special, nonminimal in terms of state dimension, parameterisations of linear dynamic systems have been developed. These special models allowed the formulation of stable adaptive control techniques, see e.g. Narendra and Annaswamy (1989), and Sastry and Bodson (1989). No such general framework exists for nonlinear systems and there is a need for some kind of nonlinear dynamic model, which allows the synthesis of control input for the plant to be performed and is at the same time `manageablea enough, allowing analytical treatment and the incorporation of mechanisms ensuring stability. Already in Narendra and Parthasarathy (1990) a possibility of using dynamic neural models combining nonlinearities with dynamics for the identi"cation of nonlinear plants was demonstrated in simulations. Still, however, relatively little has been done as far as application of such models for control, and particularly for adaptive control, is concerned. Most papers present heuristic approaches, (e.g. Gupta, Rao & Nikiforuk, 1993; Sastry, Santharam & Unnikrishnan, 1994; Parlos, Chong & Atiya, 1994) and comprehensive convergence analysis of adaptive schemes is still missing. Stability of dynamic networks in control, in a nonadaptive context, has been rigourously addressed in recent publications (Suykens, De Moor & Vandewalle, 1995a,b; Suykens, Vandewalle & DeMoor 1997; Verrelst, Van Acker, Suykens, Motmans, DeMoor & Vandewalle, 1997) using an extension of H theory. While use of = H techniques raises questions about conservativeness = of the stability results, this is certainly an important contribution. Models of dynamic systems can be constructed using neural networks in di!erent ways, for example by closing a feedback loop around a static MLP network. Modelling capabilities of such dynamic structures derive from approximation properties, for static functions, of static MLP networks. The ability of Hop"eld-type recurrent neural networks to approximate dynamic systems, in continuous and discrete time, has been shown respectively in Funahashi and Nakamura (1993) and Jin, Nikiforuk and Gupta (1995). It is our belief that recurrent neural networks of this form have the potential to provide useful models for adaptive control of many nonlinear systems. Recurrent networks of the Hop"eld type have relatively simple structure and the hyperbolic tangent function tanh( ) ) has advantageous properties like smoothness, boundedness and being monotonic. Loosely speaking it is a kind of `milda nonlinearity. Usefulness of these features will become more apparent and more precisely formulated in the following sections, where they are exploited in the stability analysis. A model of the controlled system based on a recurrent network can be interpreted as a state space model, usually with a nonminimal state dimension. Use of this type of model for adaptive control of nonlinear systems seems quite natural, as "rstly, there is no general equivalence, for nonlinear systems, between state-space and input} output models while for most physical systems, statespace models based on "rst principles are a natural way of describing dynamics. Secondly, even when the input}output models exist, like for example a SISO system y(d)"F(y(d~1),2, y)#G(y(d~1),2, y)u practical utilisation of such models, when only output y but not its derivatives y(d~1),2,y5 are measured, still requires the construction of a dynamical state-space model. If a dynamic neural model can be trained, so that in some region its input}output behaviour is close to that of the unknown system, then we should be able to obtain some kind of equivalent information about the state of the unknown plant from the state of the neural network. G.J. Kulawski, M.A. Brdys& / Automatica 36 (2000) 5}22 Thus, if a controller based on such a model can be constructed, it should be able to achieve a good dynamic performance due to the information about the state of the actual plant without the explicit use of the state and parameter observer. In the method presented here, of which a detailed description is given in the following section, a linearizing feedback control law is computed analytically for the network and applied to the plant while parameters of the network are updated on-line. Such approach can be classi"ed as indirect adaptive control, as the parameters of the plant model are adapted and control is computed based on the current model, rather than directly adapting the controller parameters. The reason for such a choice is that, as opposed to linear systems, the question of existence of direct dynamic controllers for nonlinear systems is a very di$cult one and the use of a heuristicly chosen direct adaptive dynamic controller, as for example in Ku and Lee (1992), renders the overall scheme extremely di$cult for analytical treatment. A similar general motivation appears to underlie recent schemes of Jin, Nikiforuk and Gupta (1993, 1994) and Delgado, Kambhapati and Warwick (1995), in which, either continuous or discrete-time recurrent networks, respectively, are used as a model of the unknown system and the control law utilises the state of the network. The scheme presented here was "rst proposed in Kulawski and BrdysH (1994) and BrdysH , Kulawski and Quevedo (1996) and a convergence result for the case of constant (or slowly varying) reference output was presented in BrdysH , Kulawski and Quevedo (1998) using the singular perturbation methodology. In this paper, the stability analysis is extended to the case of a general reference signal employing a completely di!erent approach. The su$cient conditions for stability are derived for exponentially stable nonlinear plants. In order to validate limitations of the assumptions made during the stability analysis, among which the assumption about perfect parameterisation is the most severe one, a comprehensive simulation study is performed. The simulation examples are selected accordingly starting from a simple academic one and ending with the induction motor case study. 2. Control algorithm We seek to design a control law for a continuous-time nonlinear system, which is at this point formulated quite generally as q5 "f (q, u), y"h(q), (1) where q3Rs is a state vector, u3Rm is an input vector, and y3Rm is an output vector. The objective is to make 7 Fig. 1. Diagonal dynamic network. Fig. 2. A single dynamic neuron. the system outputs track a vector of speci"ed trajectories y 3Rm. 3%& A recurrent neural network is used as a dynamic model of the system (1), based on which the control law is synthesised. The neural model is de"ned by x5( "Dx( #A) T(x( )#Bu, y( "Cx( , (2) where x( 3Rn is a state vector, u3Rm is an input vector, and y( 3Rm is an output vector. A nonlinear operator T( ) ) is de"ned as C D tanh(x( ) 1 T(x( )" F , tanh(x( ) n D"diag(d ,2, d ) is a n]n diagonal matrix with nega1 n tive entries, d (0, AK "diag(a( ,2, a( ) is an n]n diagi 1 n onal matrix, B3RnCm and C3RmCn. An example of such a network with two inputs and two outputs is shown in Fig. 1 and a single neuron in Fig. 2. The network can be regarded as a parsimonious version of the Hop"eld-type network. It has a diagonal structure, that is, there is no interaction between dynamics of di!erent neurons (these are shown as dashed lines in both "gures). From now on, a( will denote a vector of parameters containing the diagonal elements of AK : CD a( 1 a( " F . a( n G.J. Kulawski, M.A. Brdys& / Automatica 36 (2000) 5}22 8 The following assumption guaranteeing exponential stability of the neural model (as shown in Lemma 3 in BrdysH et al., 1998) is required. Assumption 1. a( (!d , i"1,2, n. i i P 2.1. Control synthesis The evolution of the network output can be expressed as y(5 "Cx(5 "CDx( #CAK T(x( )#CBu. (3) An assumption is made: Assumption 2. CB is invertible. The above assumption implies relative degree of the neural model equal to one. The feedback linearizing control can be calculated as u"(CB)~1(!CDx( !CA) T(x( )#v) (4) which applied to the network results in it being decoupled and linear with respect to the new input vector v. y5( "v. (5) The auxiliary control input v is designed as a simple linear pole placement v"y5 !a(y( !y ), (6) 3%& 3%& where a is a positive constant, a'0. The control input, as de"ned by (4) and (6), is applied to both the plant and the neural model. The error between the output vector of the model and the reference output vector, e( "y( !y , 3%& obeys e5( #ae( "0 and converges exponentially to the origin with the rate a. Thus, we can assume, that after some initial transient, e( converges to zero and so the plant tracking error e "y !y is equivalent to the modelling error 5 3%& e "y( !y: . e "y !y"y( !y"e . (7) 5 3%& . From now onwards, for simplicity of exposition, a single-input }single-output plant is considered. 2.2. On-line parameter adaptation On-line updates of the parameter vector a( are performed in discrete-time instants, with a period ¹, in the direction opposite to the gradient of the following error criterion: 1 E" 2 P kT`T (y( !y)2 dq, kT LE a( (k¹#¹)"a( (k¹)!j , i i La( i where j'0 is the learning rate. The discrete-time instants, when updates take place, are indexed by the positive integer k. Calculating the error function derivative as (8) (9) (k`1)T LE Ly( (y( !y) " dq (10) La( La( kT i i this derivative, LE/La( , is obtained by integrating over the i time ¹ the following two di!erential equations: d LE Ly( "(y( !y) , dt La( La( i i Lx( Ly( "c i, iLa( La( i i d Lx( Lx( i"(d #a( tanh@(x( )) i#tanh(x( ). i i i La( i dt La( i i starting with zero initial conditions: (11) (12) (13) LE Lx( i(k¹)"0. (k¹)"0, La( La( i i Once the parameter updates are done, initial conditions in Eqs. (11)}(13) are reset to zero and the cycle is repeated. Di!erential equations (11)}(13) need to be integrated in real time. Eq. (13) is obtained by applying a partial derivative L/La( to both sides of the di!erential state i equation (2) (under standard smoothness conditions we can assume that d/dt and L/La( commute). Because of the i diagonal structure of the network state equation we have Lx( /La( "0 for jOi. Eq. (13) describes the so-called sensij i tivity model, as in e.g. Narendra and Parthasarathy, (1991). Its derivation relies on the assumption that a( is i constant. As an alternative to Eq. (9), parameter updates could be performed continuously with a very small learning rate so that this assumption would be `almosta true (see Narendra & Parthasarathy, 1991). We chose the periodic update method as it is more conceptually clear and makes the overall scheme more tractable analytically. In the adaptive literature, such schemes are referred to as hybrid adaptive control. It is pointed out in Narendra and Annaswamy (1989) that they exhibit better robustness, with respect to disturbances, compared with continuous parameter adaptation. Although results in Narendra and Annaswamy (1989) refer to the control of linear systems, similar properties can be expected here as an additional nonlinearity of the tracking error due to nonlinear plant dynamics should not change the underlying argument. The updating process has to be constrained in the way that Assumption 1 is satis"ed at any point in time. The `rawa values of weights a( being the result i of Eq. (9) are projected onto the nearest point of the set: P ,Ma( : a( 4!d !eN, i i i i (14) G.J. Kulawski, M.A. Brdys& / Automatica 36 (2000) 5}22 Fig. 3. Adaptive control using recurrent network. where e is a small positive number used to construct a closed set which is needed for a well-de"ned projection. A diagram of the overall control scheme is shown in Fig. 3. 3. Stability To proceed with the convergence analysis it is assumed that there exists a choice of network parameters a( "a, such that a recurrent neural network is capable of exactly modelling the plant. In other words, the dynamical system x5 "Dx#AT(x)#Bu, 9 cal tractability and insight into the convergence issues. As a trade-o!, it may not have the full approximation capabilities of the Hop"eld network, although this could probably be compensated to some extent by a further increase of the state vector dimension. Simulations reported here seem to justify this conjecture. In any case, whether with the diagonal network (2) or a fully interconnected one, some residual structural error will remain. By the structural error it is meant that there is in fact no ideal set of parameters of Eq. (15) which would give a perfect model of Eq. (1). Although the assumption about perfect modelling capabilities made here is strong, it is not completely unreasonable. For example, many results in adaptive control assume a linear plant, although a perfectly linear plant is hardly ever the case. The stability analysis which follows provides understanding of the general mechanisms governing the behaviour of such neural adaptive systems and explains interactions between plant and model dynamics on the one hand and the parameter adaptation on the other. Incorporation of the structural error into the analysis needs certainly to be the next step in this development. To enable the controller to handle signi"cant structural errors, modi"cations of the adaptation laws will most likely be necessary. The following assumptions are required: Assumption 3. Matrix M"(D!(1/CB)BCD!(a/CB)BC) is Hurwitz (i.e. all its eigenvalues have negative real parts). (15) y"Cx is a realisation of the same input}output mapping u(t)Py(t) as Eq. (1). System (15) has its state vector of di!erent size than Eq. (1), always larger even if D and A were full matrices (see results in Funahashi and Nakamura, 1993). Hence, even if a( "a, vector x is not the same as q. The increase in the state dimension is the price to be paid for the structure of Eq. (15). This can be seen as an analogy with nonminimal parameterisations of linear dynamic systems for the purpose of adaptive control, as mentioned in the Introduction. We require that the plant (15) satis"es Assumption 1, implying exponential stability of the original plant (1). The input}output mapping realised by Eqs. (1) and (15) is certainly parameterised by the initial conditions of their respective state vectors. However, since in this analysis the plant is assumed to be stable, the in#uence of initial conditions should die out and Eq. (15) will be considered as a valid description of the controlled plant. From now on, the vector x will denote the state of the plant. The ability of Hop"eld type networks to approximate dynamic systems has already been mentioned. The simpler network structure is chosen here to allow analyti- Assumption 4. Reference output signal y and its deriva3%& tive y5 are bounded, i.e. there exist positive constants 3%& d , d such that Dy5 D(d , Dy D(d . 1 2 3%& 1 3%& 2 Although Assumption 4 prevents theoretical step inputs to be considered, it does not pose signi"cant restrictions in practice where pre-"ltering of sharp changes of reference inputs is a common approach. In addition to Assumptions 1}4 an extra Assumption 5 is needed and two conditions expressed by inequalities (27) and (29) must hold. These will be discussed later in the proof where it can be done in a much more natural way. The main result of the paper can now be stated as follows. Theorem. If the initial parameter mismatch DDa( !aDD is small enough, and ¹ long enough, control (4) achieves ultimate boundedness of all signals in both the plant (15) and the model (2). There exists a positive number jM such that for 0(j4jM the updating scheme (8)}(10) results in the bound on the tracking error e "y !y decreasing t 3%& monotonously. G.J. Kulawski, M.A. Brdys& / Automatica 36 (2000) 5}22 10 K A K BK K Proof. Part 1. Ultimate boundedness of DDx( DD, DuD, DDxDD: Because control input (4) utilises the state of the network, its dynamics in the closed loop are modi"ed to 42 x( TP NA) T(x( )#B 1 a y5 #B y CB 3%& CB 3%& x5( "Dx( #AK T(x( ) 42DDPx( DD NA) T(x( )#B 1 a y5 #B y CB 3%& CB 3%& #B 1 (!CDx( !CAK T(x( )#y5 !aCx( #ay ) 3%& 3%& CB A B A B 1 a 1 " D! BCD! BC x( # I! BC AK T(x( ) CB CB CB 1 a # By5 # By . CB 3%& CB 3%& (16) The closed-loop neural model (16) is driven only by y and y5 which are bounded. There is no direct feed3%& 3%& back from the state of the plant into Eq. (16). A feedback from the plant output into Eq. (16) is applied indirectly via the parameter updates in discrete moments of time. Therefore, we can "rst show the ultimate boundedness of DDx( DD in Eq. (16), and then proceed with the boundedness of DDxDD. We denote A B A B 1 a 1 M" D! BCD! BC , N" I! BC . CB CB CB 42DDPDD DDx( DD(DDNA) T(x( )DD 2 1 a DDBDDDy5 D# DDBDDDy D # 3%& 3%& CB CB A B 4m6 DDx( DD. This gives <QK 4!DDx( DD2#m6 DDx( DD, where A C D a( tanh(x( ) 1 1 F A) T(x( )" <K "x( TPx( , DDA) T(x( )DD4DDa( DD. where P is a symmetric positive-de"nite matrix satisfying the Lyapunov equation Therefore, <QK "x(5 TPx( #x( TPx(5 Furthermore, relation 1 1 # y5 BTPx( #x( TPB y5 CB 3%& CB 3%& a a y BTPx( #x( TPB y # CB 3%& CB 3%& "!DDx( DD2#2x( TPNA) T(x( ) a 1 y5 #2x( TPB y CB 3%& CB 3%& A "!DDx( DD2#2x( TP NA) T(x( )#B B 1 a y5 #B y . CB 3%& CB 3%& We have A 2x( TP NA) T(x( )#B DDNA) T(x( )DD4DDNDD DDA) T(x( )DD4DDNDD DDa( DD. 2 2 If we de"ne m"max a( m6 , then Eq. (17) holds with the @@ @@ same m for all t. Updating of a( , as will be described later, ensures that DDa( DD remains bounded. Since all other quantities appearing in Eq. (18) are bounded, m is a bounded positive number. Therefore <QK (0 ∀DDx( DD'm. "x( T(MTP#PM)x( #T(x)TA) TNPx( #x( TPNA) T(x( ) #2x( TPB (18) The above expression for m6 is obtained by noticing that Consider the Lyapunov function for Eq. (16): The existence of such a matrix is assured by Assumption 3. Taking the time derivative of the Lyapunov function gives B 1 a m6 "2DDPDD DDNDD DDa( DD# DDBDDd # DDBDDd . 2 2 1 CB 2 CB a( tanh(x( ) n n and due to the fact that Dtanh(x)D(1 MTP#PM"!I. (17) B 1 a y5 #B y CB 3%& CB 3%& !DDx( DD2#mDDx( DD4!k DDx( DD2, 1 where 0(k (1, is satis"ed for 1 m DDx( DD5 . 1!k 1 Therefore, outside any ball with radius greater or equal to m "m/(1!k ), the right-hand side of Eq. (17) can be 1 1 bounded by <QK 4!k DDx( DD2 ∀DDx( DD5m . 1 1 From the above and the fact that Lyapunov function <K satis"es j (P)DDx( DD24<K (x( )4j (P)DDx( DD2 .*/ .!9 G.J. Kulawski, M.A. Brdys& / Automatica 36 (2000) 5}22 one can conclude, based on Theorem 4.10 in Khalil (1992), that there exists a "nite time t and positive 1 constants k ,c '0 such that 2 2 DDx( (t)DD4k DDx( (t )DDexp(!c (t!t )), ∀t 4t(t , (19) 2 0 2 0 0 1 S j (P) .!9 m ∀t5t . (20) 1 j (P) 1 .*/ In other words, the norm of x( decreases exponentially and in "nite time t enters a ball of radius ox( . Naturally, 1 boundedness of x( implies boundedness of y( . Assumption 4, relations (19), (20) and boundedness of DDa( DD imply that the input to both plant and model, as given by Eq. (4), also converges in "nite time to bounded values. There exists a positive constant o '0, such that u DuD4o ∀t5t . u 1 We proceed to show that x is bounded as well. A Lyapunov function for the plant (15) is chosen as DDx( (t)DD4ox( " <"1xTx. 2 Its time derivative is given by <Q "xTDx#xTAT(x)#xTBu. Since u is bounded for t5t , T(x) contains bounded 1 elements and norms of A and B are bounded, it holds that xT(AT(x)#Bu)4f DDxDD ∀t5t , 1 where f is a bounded positive constant. Thus, with D being a diagonal matrix, we have <Q 4!min Dd DDDxDD2#f DDxDD ∀t5t , i 1 i and following a similar line of thought it can be concluded that the Lyapunov derivative satis"es f ∀t5t , <Q 4!k DDxDD2 ∀DDxDD5 1 3 min Dd D!k i i 3 where k is a positive constant satisfying 0(k ( 3 3 min Dd D. Thus, assumptions of Theorem 4.10 in Khalil i i (1992) are satis"ed for t5t . It can be then concluded, 1 similarly as in the case of convergence of the norm of x( , that the norm of x decreases exponentially and in "nite time enters a ball of radius ox DDxDD4ox ∀t5t , 2 where ox is some positive constant and t 5t . This in 2 1 turn results in bounded y for t5t 5t . 2 1 Thus, ultimate boundedness of all signals in the plant and neural model has been established, that is, x, x( , y, y( and u are bounded for t5t . 2 Part 2. Convergence of the tracking error: We proceed to show how parameter updates result in a decrease of the tracking error. 11 Control input (4) makes the output of the neural model converge quickly to the reference trajectory (see relation (7)). Thus, after some fast transient when the algorithm is initialised, the tracking error is directly related to the state error e "y !y"y( !y"C(x( !x)"Cx8 . t 3%& Therefore "rstly, the Lyapunov function: <I "1x8 Tx8 2 is used, to show that the state error norm x8 converges to a ball whose size is determined by the parameter error norm DDa( !aDD"DDa8 DD. Denoting x8 "x( !x and a8 "a( !a , from Eqs. (2) i i i i i i and (15) we have x58 "d x( !d x #a( tanh(x( )!a tanh(x ) i i i i i i i i i "d x8 #a( tanh(x( )!a tanh(x ). i i i i i i Adding and subtracting a( tanh(x ) to the right-hand side i i of the above equation we obtain x58 "d x8 #a( tanh(x( )!a( tanh(x ) i i i i i i i #a( tanh(x )!a tanh(x ) i i i i "d x8 #a( (tanh(x( )!tanh(x ))#a8 tanh(x ). i i i i i i i The Lyapunov derivative is given by n <QI "x8 Tx85 " + (d x8 2#a( (tanh(x( ) i i i i i/1 ! tanh(x ))x8 #a8 tanh(x )x8 ) i i i i i n " + (d Dx8 DDx8 D#a( Dtanh(x( ) i i i i i i/1 ! tanh(x )DDx8 D#a8 tanh(x )x8 ) i i i i i n 4 + (d Dx8 DDx8 D#a( Dtanh(x( ) i i i i i i/1 ! tanh(x )DDx8 D#Da8 DDtanh(x )DDx8 D) i i i i i n a( " + Dd DDx8 D !Dx8 D# i Dtanh(x( ) i i i i Dd D i i/1 Da8 D ! tanh(x )D# i Dtanh(x )D . i i Dd D i The following relations hold: Firstly, due to the projection algorithm used (14), the following is satis"ed uniformly in time: A a( 4!d !e, i i a( e i 41! , !d !d i i a( e i 41! . Dd D Dd D i i B G.J. Kulawski, M.A. Brdys& / Automatica 36 (2000) 5}22 12 De"ning l "1!e/Dd D if 1!e/Dd D'0 and l "0 otheri i i i wise, we obtain a( i 4l (1, i"1,2, n. i Dd D i Secondly, function tanh( ) ) satis"es DDx8 (k¹)DD4o DDa8 ((k!1)¹)DD, (26) 1 where o "1/((1!max l )min Dd D!k ). 1 i i i i 4 It remains to be shown that parameter updates result in the decrease of the parameter error magnitude DDa8 DD. This is done by showing that, on the system trajectories in discrete-time instants a( (k¹), x( (k¹), the tracking error criterion E as a function of a( (k¹) and x( (k¹) satis"es k conditions (A.1)}(A.5) necessary for Lemma 1 to hold (see Appendix A). We can treat both the output of the plant and the output of the neural model as being generated by the same operator Y( ) ) whose arguments are the initial conditions of state vectors, parameter sets and input over a time interval Dtanh(x )D(1, i Dtanh(x( )! tanh(x )D4Dx( !x D"Dx8 D, i"1,2, n. i i i i i Using the above one obtains A A B Da8 D n <QI 4 + Dd DDx8 D !Dx8 D#l Dx8 D# i Dtanh(x )D i i i i i i Dd D i i/1 n Da8 D " + Dd DDx8 D !(1!l )Dx8 D# i i i i i Dd D i i/1 n n 4 + !(1!l )Dd DDx8 D2# + Da8 DDx8 D. i i i i i i/1 i/1 Finally, we obtain A B B (21) <QI 4! 1!max l min Dd DDDx8 DD2#DDa8 DD DDx8 DD. i i i i For a constant a( (k¹), i.e. between consecutive parameter updates, we can apply similar reasoning as in the "rst part of the proof to show the following. Outside the ball G ball B x8 . As a consequence, after each update, the state o (k) error enters the corresponding ball before the next update is done. Thus, from Eqs. (25) and (22) it can be concluded that, at the start of a new integration period, we have H 1 x8 :DDx8 DD(m " DDa8 (k¹)DD 2 (1!max l )min Dd D!k i i i i 4 (22) the Lyapunov derivative (21) can be bounded by <QI 4!k DDx8 DD2 ∀DDx8 DD5m (a( (k¹)), (23) 4 2 where k is a positive constant satisfying 4 0(k ((1!max l )min Dd D. Using again Theorem 4.10 4 i i i i from Khalil (1992), it can be concluded that DDx8 (t)DD4DDx8 (k¹)DDexp(!c t) ∀k¹4t(k¹#q , (24) 4 k DDx8 (t)DD4ox8 (k)"m (k) ∀k¹#q 4t4(k#1)¹, (25) 2 k where c is a positive constant. When a( (k) is constant, 4 state error norm enters in "nite time q a ball, whose size k ox8 (k) is determined by the magnitude of the parameter error DDa8 DD. Constant c is uniform on k, due to the fact that 4 the Lyapunov function <I does not depend on time and the bound on its derivative (23) is uniform. Therefore, if during updates, changes in DDa8 (k)DD are uniformly bounded and thus changes in ox8 (k) are uniformly bounded, q is k uniformly bounded. In other words, there exists a time q6 such that q 4q6 ∀k'0. k We need to choose the update period as ¹'q6, that is, longer than the time needed by the state error to enter the y(t)"Y(x(k¹), a, u ), *kT, t+ y( (t)"Y(x( (k¹), a( (k¹), u ) for k¹4t4(k#1)¹. *kT, t+ The tracking error criterion for each integration period P (k`1)T (y( (q)!y(q))2 dq kT is a function of a( (k) and x( (k) and it has a global minimum E ,0 in a( (k)"a, x( (k)"x(k). Let us assume that k for nonzero input E (a( (k), x( (k))'E (a, x(k)) for k k [a( (k)T, x( (k)T]O[aT,x(k)T]. Since tanh( ) ) is a smooth function, the function on the right-hand side of the state equation of the neural model (2) is smooth with respect to both x( and a( . As a consequence, solutions of state equations are smooth with respect to the initial conditions x( (k) and parameters a( (k). In other words, Y( ) ) is a smooth operator. Thus, E is a smooth function of a( (k) k and x( (k). As such it satis"es Eq. (A.2) and, in a "nite ball around the minimum, also Eq. (A.4). 1 E" k 2 Assumption 5. We have to assume that the Hessian of the error function is positive dexnite in a neighbourhood of the minimum (but not necessarily in the optimal point itself, where the updates are ewectively not performed). For this to hold it is su$cient, but not necessary, that the error function E satis"es the su$cient conditions for k the existence of a local minimum in [aT, x(k)T]. We also assume that on the system trajectories the following is satis"ed: K K LE(a( (k), x( (k)) 4cDDa8 (k)DD ∀k, La( (27) G.J. Kulawski, M.A. Brdys& / Automatica 36 (2000) 5}22 13 where c is a positive constant. It can then be shown (see Appendix B) that conditions (A.3) and(A.5) are satis"ed if the update step j satis"es Due to Eq. (25), the above relation also implies decrease of the tracking error bound. h jQ 4jHH The analysis presented in the proof of the theorem highlights mechanisms governing stability and tracking error convergence in this neural adaptive control system. Firstly, with bounded reference input and with constant values of the parameters the tracking error remains bounded. Then, decrease of the tracking error is achieved in the updating process via decrease of the network parameters distance to the `truea parameters of the plant, that is, the parameter values which result in a perfect modelling of the plant by the recurrent network. Consequently, two types of dynamics can be distinguished in this system: the fast continuous-time dynamics of the plant and neural controller and the slow discrete-time dynamics of learning. Learning dynamics binds together the evolution of parameter error and state error from one update step to another. The speed of the plant and neural model dynamics determines how fast the learning process can progress. This dependence is quanti"ed by the length of the update period ¹, which must be long enough so that relation (26) holds and the gradient used gives a direction of improvement of the parameter error. The result of the theorem does not o!er prescriptive methods for choosing ¹ and j and in practice, as it is the case in the reported simulations, they have to be chosen experimentally. It provides however an insight into the convergence mechanisms which helps to guide such process. 1!o max[(1#gj /j )DDL2E(a,x6 )/La( Lx( DD /(j (1!g))] 1 .*/ .!9 2 .*/ " c (28) and if the following holds: (1#gj /j )DDL2E(a, x6 )/La( Lx( DD .*/ .!9 2(1, (29) j (1!g) .*/ where j , j are the smallest and greatest eigenvalues .*/ .!9 of the matrix of second derivatives of E with respect to a( . The maximum of the expression in the fraction in Eq. (28) is taken over the ball in which system trajectories are contained. Bounded initial parameter mismatch, results of the "rst part of this proof and the following analysis show that system trajectories are indeed restricted to a ball. It is very di$cult to specify precise conditions such that relation (29) is always satis"ed. However, certain mechanisms of the state error and parameter error convergence appear quite clearly. The presence of the initial conditions error x8 (k) acts as a kind of disturbance in the updating process, since it perturbs the minimum of E form the point of zero parameter error a( (k)"a. This k disturbance x8 (k) has to be small enough relative to the magnitude of the parameter error a8 (k¹), so that updates in the direction of negative gradient LE/La( can improve the parameter error. The shape of the error criterion function can be in#uenced by varying the length of the integration period ¹. It appears that the fraction appearing on the left-hand side of relation (29) can be reduced by increasing the length of the integration period ¹. Due to the stability of both plant and neural model, the in#uence of the initial conditions on the outputs of plant and model, and thus on tracking error, decreases with time. Thus, increasing the integration period ¹ reduces the in#uence of the initial conditions mismatch x8 (k) on the error criterion relative to the in#uence of the parameter mismatch a8 (k¹). We can therefore expect that the ratio o max 1 DDL2E(a, x6 )/La( Lx( DD 2 j (L2E(a6 , x( )/La6 (k)2) .*/ is smaller for bigger ¹. With all the conditions (A.2)}(A.5) satis"ed, Lemma 1 can be called to show that there exists a choice of learning rate 0(j(jH such that the parameter error decreases in the consecutive steps. To preserve condition (28), the maximum value of the learning rate needs to be chosen as jM "minMjH, jHHN. Thus for 0(j(jM it holds that DDa8 (k#1)DD2(DDa8 (k)DD2 ∀k'0. 4. Simulation examples Example 1. This is a rather academic example in which the controlled plant is chosen as a recurrent network of the structure (15). Thus, the assumption about the recurrent network being able to perfectly model the plant for a certain set of network parameters is satis"ed. This allows to verify theoretical results presented earlier. Both the object-network and the controller-network have three states. There is initially a parameter mismatch between a( 's in the controller-network and actual values i of a 's. Only a( 's are updated using the procedure dei i scribed in Section 2, while the rest of the parameters, i.e. elements of D, B, C are assumed to be known. The update step is ¹"0.5, the learning rate initially j"1 is later increased to j"200. Convergence of the plant output to the reference output is shown in Fig. 4. Fig. 5 illustrates that learning results in both model parameters and model states converging to the actual plant parameters and states, respectively. Simulations with all the parameters being updated showed that tracking error convergence does not necessarily mean that parameter and state errors need to converge. That appears dependent on the character of the reference output which bears 14 G.J. Kulawski, M.A. Brdys& / Automatica 36 (2000) 5}22 Fig. 4. Control of a neural network. Reference output (dotted line) and plant output (solid line). Fig. 5. Control of a neural network. (a) parameter errors convergence: a8 (solid line), a8 (dashed), a8 (dot dash); (b) state errors convergence: x8 1 2 3 1 (solid), x8 (dashed), x8 (dot dash). 2 3 resemblance to the problem of persistency of excitation known in adaptive control. Example 2. The control object is a single robot arm described by fQ "f , 1 2 fQ "!a sin(f )!a f #bu, 2 1 1 2 1 where y"f is the arm position which is the measured 1 output, f is the (unmeasured) angular velocity and input 2 u is the torque. The state dimension of the neural model was chosen equal to 10, after a few trials with di!erent network sizes. Elements of D and A) were initialised as Fig. 6. Control of a single arm. (a) y (dotted) and angular position of 3%& the arm (solid), (b) control input. random numbers within the interval [!1,0] with uniform distribution, elements of B, C were similarly initialised in the interval [0,1]. Prior to its application in on-line control the network was trained o!-line to obtain a "rst approximation of the plant model. This relates to the fact that the theorem requires the initial parameter mismatch between parameters of the neural model and plant to be appropriately small. Thus, the o!-line training phase can be understood as providing initial model parameters which are close enough to the ideal ones. At this stage all parameters, i.e. D, A) , B, C, were updated using the same methodology as for updates of A) described in Section 2.2. To ensure that parameters d rei main negative, the projection algorithm was extended to constrain each d to the set d 4!e. An input signal i i consisting of two sine waves, u(t)"4 sin(0.5t)# 0.5 sin(2t), was applied to both the network and the plant and training was carried on until errors between the outputs of the network and the plant were small. Prior to on-line control satisfaction of Assumption 2 was checked. In both the o!-line training and control e"0.0001 was used. As opposed to Example 1, there is an inevitable structural error in the neural model. Validation of the controller robustness with respect to this error is one of the major objectives in this simulation example. Control is performed according to the algorithm described in Section 2. Results for two di!erent reference trajectories are shown in Figs. 6 and 7. These trajectories are obtained by passing a piecewise constant and a triangular signal, respectively, through a "rst-order stable linear "lter. In both simulations ¹"0.2 and j"100. After initial #uctuations, the output of the plant converges to the reference trajectory. The increased output error seen in Fig. 6 at time t"50 is due to the mass of the G.J. Kulawski, M.A. Brdys& / Automatica 36 (2000) 5}22 15 Fig. 8. MIMO control of the induction motor using dynamic networks. Fig. 7. Control of a single arm. (a) y (dotted) and angular position of 3%& the arm (solid), (b) control input. General arguments for the use of dynamic neural models for nonlinear adaptive control of systems with state, or part of state, unavailable for measurement have been already given in the Introduction. Clearly, induction motor falls into such category and is in fact one of the case studies motivating the development of the scheme proposed in Section 2. 5.1. MIMO control of the motor link being increased by 50%. The increased output error in Fig. 7 at time t"55 is a result of the mass being reduced by 50%. In both cases, the controller is able to adapt to the change and regain good tracking performance. 5. Induction motor control In this section, a multi-input}multi-output control scheme for the induction motor is presented which is based on the neural algorithm described in Section 2. Induction motor control is a di$cult, and still not completely resolved, engineering problem. This is due to highly nonlinear dynamics of the machine and unavailability of measurements of some of the state variables and usually also of some of the controlled outputs, in a typical hardware con"guration. In addition to that, some of the machine parameters, mainly the rotor resistance, exhibit strong variations due to changing thermal conditions. The state-space description of the motor (C.1)}(C.5) uses a state vector [u , i , i , t , t ] consisting of rotor 3 4$ 42 3$ 32 speed, stator current components and rotor #ux components (see Appendix C). Only stator currents and rotor speed are normally available for measurement. In some control schemes for the induction motor, observers are used to obtain estimates of the unmeasurable variables. However, as the plant is nonlinear, the separation principle does not apply and stable state feedback control combined with a stable observer do not imply stability of the closed-loop system. This is further complicated by signi"cant parameter variations during operation. Due to these di$culties, no stability proof for observer-based induction motor control schemes has been shown so far. A general diagram of the proposed control structure is shown in Fig. 8. Rotor speed u and amplitude of the 3 stator #ux Dw D are the controlled variables. Two inputs 4 generated by the controller are the amplitude Du D and the 4 angular frequency of stator supply voltage. Amplitude and frequency of the required sinewave are typically the inputs of a voltage}source inverter. The controller requires an estimate of the stator #ux magnitude Dw D, which can be obtained from a stator #ux 4 observer. It has to pointed out that, "rstly, it is easier to estimate stator than rotor #ux, and secondly, a need for the magnitude only is a weaker requirement than relying on estimates of the full stator phasor. Compared to methods which require good estimates of both coordinates of the stator phasor for conversion between di!erent reference frames, the sensitivity of the method presented here with respect to observer errors will be much lower. Furthermore, as the updating scheme of the neural algorithm applied here is based on the integral of the output error, this will further provide for averaging out the estimation errors. 5.2. Dynamic neural controller for the induction motor The control algorithm utilises the technique presented in Section 2, with a modi"cation in the structure of the neural model of the plant. A dynamic network serves as a model of the motor, with stator voltage magnitude Du D 4 and angular supply frequency u being the inputs and 4 stator #ux amplitude Dw D and rotor speed u as the 4 3 outputs. The neural model of the motor is constructed as x5( "Dx( #AT(x( )#f tanh(i )#f tanh(i )#Bu, 1 4x 2 4y y( "Cx( , (30) 16 G.J. Kulawski, M.A. Brdys& / Automatica 36 (2000) 5}22 where, as before, x( 3Rn is the state vector, u"[Du D,u ]T is 4 4 the input vector and y( "[DwK D,u( ]T is the output vector. 4 3 f , f 3Rn are vectors of adjustable weights. i and 1 2 4x i denote components of the stator currents phasor with 4y respect to a rotating reference frame "xed to the stator input voltage phasor u . Measurements of stator currents 4 are easily obtainable and, since u is known, the conver4 sion from (sD, sQ) coordinates can be easily performed. The extra terms, containing measurements of stator currents, were found to improve the modelling capabilities of the network and consequently also the quality of control. As already mentioned, the proposed structure of the dynamic network (2) is a compromise between the full modelling capabilities and analytical tractability of the algorithm. It is shown here that this structure can be augmented by utilising the available information and thus improve the modelling capabilities at no extra cost. Since in normal operation, the value of the rotor speed expressed in rad/s is about two orders of magnitude bigger than the #ux amplitude expressed in Wb, the speed output, that the neural model is supposed to model, is the measured value of rotor speed in rad/s scaled down by a factor of 0.01. This is done to improve conditioning of the error function used in parameter adaptation. Consequently, the speed setpoint for the motor (in rad/s) is also scaled by 0.01 for control generation purposes. Following the general procedure described in Section 2, control input is calculated as u"(CB)~1(!CDx( !CAT(x( )!Cf tanh(i ) 1 4x !Cf tanh(i )#v), 2 4y where, as previously, (31) v"y5 !a(y( !y ), 3%& 3%& and y contains the reference values for #ux magnitude 3%& and speed, y "[Dw DH, uH]T. 3 3%& 4 Although the presence of the stator current measurements in the model (30) resembles an observer structure, it has to be pointed out that, "rstly, as opposed to typical observers based for example on the extended Luenberger scheme (BrdysH & Du, 1991; Du & BrdysH , 1993), here measurements enter in a nonlinear fashion and there is no explicit corrector term. Secondly, the control input is generated based on the model (30) in an integrated fashion without use of the explicit state and parameter observer and application of the certainty equivalence principle. form distributions, for D, A in the interval [!1,0], for B, C in [!0.01,0.01] and f , f in [!0.05,0.05]. The 1 2 network was "rst trained in open loop with externally generated control inputs applied to both motor and neural model. All the weights of the network were updated in this phase. The projection algorithm extended to D, as described in Example 2, was used both for the preliminary training and the control. e"0.0001 was used. Similarly as before the size of the network and the initialisation intervals were chosen after a few trials. In the control experiment, the start-up of the motor was performed using an independent controller whose inputs were applied both to the motor and the neural model. In this start-up period learning was switched o!. After the start-up, the control was switched to the neural controller described above. During the closed-loop control, only weight matrices D, A, f and f were adapted, 1 2 with period ¹"0.2 s and learning rate j"10. Stator #ux magnitude was obtained directly from the simulation model without implementing the observer. In the experiment, the motor is subjected to a constant load torque t "5 N m and it is expected to follow a traL jectory of reference rotor speed. This is a standard test trajectory consisting of an increasing (acceleration) and then decreasing (deceleration) ramp, which is passed through a stable linear "lter. The reference value of the stator #ux magnitude is kept constant at Dw DH"1.1 Wb. 4 Results are shown in Fig. 9. The controller is able to follow the speed pro"le quite well while maintaining the #ux amplitude close to the desired value. The behaviour of four randomly chosen weights of the neural model, during the course of this experiment, is shown in Fig. 10. It can be seen that speed variation over a wide range necessitates adaptation of the neural model, whose approximation capabilities are not global enough. 5.3. Simulation results For the simulation study reported here, a network of the structure (30) with 40 dynamic neurons was used. It was "rst trained in open-loop with externally generated control inputs applied to both motor and neural model. Weights were initialised as random numbers with uni- Fig. 9. Speed pro"le following. (a) reference speed (dotted) and rotor speed (solid) (rad/s), (b) stator #ux magnitude (solid) and reference value (dotted) (Wb). G.J. Kulawski, M.A. Brdys& / Automatica 36 (2000) 5}22 17 Fig. 10. Speed pro"le following, changing weights in the neural model: (a) weight a , (b) weight d , (c) weight f , (d) weight f . 2 1 1,2 2,19 Remark. As already mentioned, the analysis of the structural error is beyond the scope of the theoretical work presented in this paper. Control of the induction motor presented above is clearly a case where structural error is present. The recurrent network is not able to provide a global model of the complex dynamics of the motor and parameter adjustments are required if the operating point of the machine changes signi"cantly. In this way parameter adaptation appears to give robustness with respect to the structural error. The issue of the structural error is related to the locality of the neural model and the question whether the network will `forgeta a model obtained at a previous operating point. In the convergence analysis, which is presented for the case of no structural error, parameter updates result in the decrease of both the parameter error and the tracking error bound. Thus, the overall quality of the (global) model improves. The structural error will prevent achieving a good global model and, as the operating point changes, the network may, to some extent, `forgeta the previous model. Intuitively however, an accurate enough local model is su$cient to generate suitable control action in the present operating region and consequently achieve maintaining decrease of the tracking error. This explains the observed robustness with respect to the structural error. 6. Conclusions The adaptive neural control algorithm for nonlinear exponentially stable plants presented here is based on use of a recurrent neural network as a dynamic model of the system. The recently obtained stability results are extended to the case of general reference output signals. The analysis presented in the proof of the theorem highlights the mechanisms governing the stability of such neural control systems, and the interactions between the dynamics of the plant and the neural model on the one hand, and the dynamics of learning on the other. Speed of the plant-model dynamics limits the admissible speed of adaptation. This is quanti"ed by the smallness of the adaptation rate and the length of update period ¹. The stability analysis shows that recurrent networks seem to possess certain intrinsic features making them suitable for nonlinear adaptive control in the case of unmeasurable state of the plant. It is clear from the presented results that hyperbolic tangent activation function appears to be a very good choice of nonlinearity for constructing nonlinear dynamic models. Properties of the function tanh( ) ), like smoothness, boundedness and being monotone, have been used many times in the stability proof in a crucial way. It is our belief that appropriately structured recurrent neural networks can provide conveniently parameterized dynamic models for many nonlinear systems for use in adaptive control. Further research e!ort needs to be directed on the one hand towards analysis of approximation capabilities of the recurrent networks and improvements in the model structure to remove the restrictive assumptions, and on the other hand towards possible improvements in learning methods and their stability as well as incorporation of the modelling errors into the analysis. G.J. Kulawski, M.A. Brdys& / Automatica 36 (2000) 5}22 18 K Appendix A. Lemma A.1 with proof 4 !2jbDDa8 (k)DD Lemma A.1. Consider a smooth function E(z( ) which has a local minimum in z( "z, i.e. E(z( )'E(z) ∀DDz( !zDD(r, z( Oz (A.1) and LE(z) "0. Lz( K (A.2) Vector variable z( is composed of two vectors of the same dimension z( "[a( T, x( T]T. Variable a( is updated according to LE(z( (k)) a( (k#1)"a( (k)!j . La( K K 0(b(1, K K L2E(z( ) 4M(R, Lz( 2 2 DDx( (k)!xDD4sDDa( (k)!aDD, s(R. LE(a( , x( ) LE(a, x( ) L2E(a6 , x( ) " # (a( !a) La( La( La( 2 L2E(a6 , x( ) LE(a, x) L2E(a, x6 ) # (x( !x)# a8 " La( Lx( La( 2 La( L2E(a, x6 ) L2E(a6 , x( ) " x8 # a8 , La( Lx( La( 2 a6 "a a( #(1!a )a, 1 1 (A.4) x6 "a x( #(1!a )x. 2 2 (A.5) Positive numbers a and a satisfy 0(a , a (1. 1 2 1 2 Thus, using Eqs. (A.4) and (A.5), then there exists a positive number jH such that for 0(j4jH the following is true for all k50: K K K K # The update equation is equivalent to LE(z( (k)) a8 (k#1)"a8 (k)!j . La( From the above and using Eq. (A.3) we obtain K L2E(a6 (k), x( (k)) DDa8 (k)DD La( 2 2 4M(s#1)DDa8 (k)DD. where a8 "a( !a. Proof. Initial condition satisfying DDa( (0)!aDD(r/Js2#1 together with relation (A.5) ensure that z( (0) is inside the ball, where all assumptions are satis"ed. K LE(z( (k)) L2E(a, x6 (k)) DDx8 (k)DD 4 La( La( Lx( 2 DDa8 (k#1)DD2(DDa8 (k)DD2, S LE(z( (k)) . La( (k) (A.3) DDa( (0)!aDD(r/Js2#1, s2r2 r2 ( # "r. s2#1 s2#1 !2bDDa8 (k)DD#j where x8 "x( !x and in the ball DDz( !zDD(r, for z( Oz, and the initial condition a( (0) satisxes DDz( (0)!zDD"JDDx( (0)!xDD2#DDa( (0)!aDD2 K KB Applying the mean value theorem twice, and utilising the fact that, due to Eq. (A.2), LE(a, x)/La( "0, the following can be obtained: The second argument x( can vary as well and x( (k) denotes the corresponding trajectory. If, for all k50, it holds that LE(z( (k))T LE(z( (k)) (a( (k)!a)5 b DDa( (k)!aDD, La( La( KA LE(z( (k)) La( "j K K K LE(z( (k)) LE(z( (k)) 2 #j2 La( La( Finally, DDa8 (k#1)DD2!DDa8 (k)DD2 K 4j K LE(z( (k)) DDa8 (k)DD(!2b#jM(s#1)) La( and for 0(j(jH"2b/M(s#1) the above expression is negative. The above and the initial condition DDz( (0)!zDD2 ensure that for 0(j(jH z( (k) remains in the ball, where the assumptions are satis"ed, for all k50. h DDa8 (k#1)DD2!DDa8 (k)DD2 " a8 (k#1)Ta8 (k#1)!a8 (k)Ta8 (k) " !2ja8 (k)T LE(z( (k)) LE(z( (k))T LE(z( (k)) #j2 La( La( La( Appendix B. Learning rate condition (28) In this section it is shown that a learning rate satisfying (28), provided that (29) holds, ensures that the error G.J. Kulawski, M.A. Brdys& / Automatica 36 (2000) 5}22 criterion function E satis"es conditions (A.3) and (A.5) on the system trajectories in the discrete-time instants a( (k), x( (k). B.1. Satisfaction of condition (A.3) We "rst show that conditions (28) and (29) imply that relation (A.3) holds for the error function E on the system trajectories. Using twice the mean value theorem and the fact that the gradient of the error function is zero at the optimum point a( (k)"a, x( (k)"x(k), we have L2E(a, x6 ) L2E(a6 , x( ) " x8 # a8 , La( Lx( La( 2 B B K B K DDx8 (k)DD. A B A B L2E(a6 (k), x( (k)) j b(k)"g .*/ La( 2 j .!9 with 0(g41, giving (1#gj /j )DDL2E(a, x6 (k))/La( Lx( DD .*/ .!9 2DDx8 (k)DD. DDa8 (k)DD5 (1!g)j .*/ (B.3) LE(a( , x( ) (a( (k)!a)T La( If the above holds, condition (A.3) is satis"ed for all k50 with a certain b(k). We assume that there exists bH such that L2E(a, x6 (k)) L2E(a6 (k), x( ) x8 (k)#a8 (k)T a8 (k) La( Lx( La( 2 K LE(a( (k), x( (k)) . La( L2E(a6 (k),x( ) L2E(a,x6 (k)) x8 (k)#a8 (k)T a8 (k) La( 2 La( Lx( B L2E(a6 (k), x( (k)) DDa8 (k)DD2 La( 2 Thus and, due to Eq. (B.1), K K A K DDa8 (k)DD5 DDa8 (k!1)DD!j LE(a( (k), x( (k)) L2E(a, x6 (k)) DDx8 (k)DD 4 La( La( Lx( 2 #j .!9 Then (B.3) being satis"ed for all k50, implies that condition (A.3) holds uniformly for all k50 with b"bH. Because of the relation (26), a bound on the state error magnitude DDx8 (k)DD is determined by parameter error magnitude in the previous time instant DDa8 (k!1)DD. Also a8 (k), through the update equation, is determined by a8 (k!1) and the update step j. We show that j satisfying (28) guarantees that (B.3) holds. The update equation is equivalent to LE(a( (k), x( (k)) a8 (k#1)"a8 (k)!j . La( L2E(a,x6 (k)) !DDa8 (k)DD DDx8 (k)DD La( Lx( K K 0(bH(b(k) ∀k (B.2) Since K K L2E(a, x6 (k)) DDx8 (k)DD La( Lx( 2 L2E(a6 (k), x( (k)) #j DDa8 (k)DD . .!9 La( 2 b can be chosen as Positive numbers a and a satisfy 0(a ,a (1. 1 2 1 2 We need A AKK A 5bDDa8 (k)DD (B.1) x6 "a x( #(1!a )x, 2 2 5j .*/ L2E(a, x6 (k)) ! DDa8 (k)DD DDx8 (k)DD La( Lx( A a6 "a a( #(1!a )a, 1 1 a8 (k)T B L2E(a, x6 (k)) La( Lx( 2 DDa8 (k)DD5 L2E(a6 (k), x( (k)) L2E(a6 (k), x( (k)) !bj j .!9 .*/ La( 2 La( 2 where K A L2E(a6 (k), x( (k)) DDa8 (k)DD2 La( 2 j .*/ (1#b) LE(a, x) L2E(a, x6 ) L2E(a6 , x( ) " # (x( !x)# a8 La( La( Lx( La( 2 5 bDDa8 (k)DD relation (B.2) is satis"ed if the following holds: This leads to the condition LE(a( , x( ) LE(a, x( ) L2E(a6 , x( ) " # (a( !a) La( La( La( 2 " a8 (k)T 19 KK LE(a( (k!1), x( (k!1)) . La( Due to Eqs. (29) and (27), jHH from Eq. (28) satis"es B L2E(a6 (k), x( (k)) DDa8 (k)DD La( 2 1 DDa8 (k)DD jHH4 4 ∀k50. c DDLE(a( (k), x( (k))/La( DD G.J. Kulawski, M.A. Brdys& / Automatica 36 (2000) 5}22 20 Condition (28) implies that j(1/c, and thus the denominator in the above expression is positive. Therefore for j satisfying Eq. (28), we have K DDa8 (k!1)DD!j K LE(a( (k!1), x( (k!1)) 50. La( Appendix C. Induction motor model Thus K DDa8 (k)DD5DDa8 (k!1)DD!j K LE(a( (k!1), x( (k!1)) . La( (B.4) The above and Eq. (26) imply, that for Eq. (B.3) to hold it is su$cient that K DDa8 (k!1)DD!j K LE(a( (k!1), x( (k!1)) La( (1#gj /j )DDL2E(a, x6 (k))/La( Lx( DD .*/ .!9 2o DDa8 (k!1)DD. 5 1 (1!g)j .*/ This inequality is satis"ed with 1!((1#gj /j )DDL2E(a, x6 )/La( Lx( DD /j (1!g))o .*/ .!9 2 .*/ 1 j4 DDLE(a( (k!1), x( (k!1))/La( DD DDa8 (k!1)DD. Therefore, using Eq. (27), we conclude that the above is satis"ed uniformly for all k50 by j given by Eq. (28), provided that Eq. (29) holds. B.2. Satisfaction of (A.5) We proceed to show that such value of j also guarantees that relation (A.5) i.e. DDx8 (k)DD4sDDa8 (k)DD holds with a constant s uniformly for k50. Eq. (B.4) implies that for the above to hold it is su$cient that A K DDx8 (k)DD4s DDa8 (k!1)DD!j KB LE(a( (k!1), x( (k!1)) . La( Since due to Eq. (27), A K DDa8 (k!1)DD!j LE(a( (k!1), x( (k!1)) La( KB 5 DDa8 (k!1)DD!jcDDa8 (k!1)DD relation (26) can be used to obtain a further su$cient condition: o DDa8 (k!1)DD4s(1!jc)DDa8 (k!1)DD 1 which is satis"ed for all k50 with o s5 1 . 1!jc A three-phase squirrel-cage induction motor is considered. A few standard simplifying assumptions are made: the air gap between rotor and stator is uniform and constant, saturation and hysteresis are neglected and the stator supply neutral point is isolated. Using the Parks transformation (e.g. Vas, 1990), the three-phase stator windings (sA, sB, sC) can be transformed into equivalent quadrature-phase windings (sD, sQ). The dynamics of the motor are then given by a "fth-order nonlinear di!erential model (Marino, Peresada & Valigi, 1993) du nM t 3" p (t i !t i )! L, (C.1) 3d 4q 3q 4d dt J¸ J 3 di MR nM M2R #¸2R 3 4i 4d" 3t # p ut ! 3 4d p¸ ¸2 dt p¸ ¸2 3d p¸ ¸ 3 3q 4 3 4 3 4 3 1 u , (C.2) # p¸ 4d 4 di nM MR M2R #¸2R 3 4i 4q"! p u t # 3t ! 3 4q p¸ ¸2 dt p¸ ¸ 3 3d p¸ ¸2 3q 4 3 4 3 4 3 1 u , (C.3) # p¸ 4q 4 dt R R 3d"! 3t !n u t # 3Mi , (C.4) p 3 3q ¸ 4d dt ¸ 3d 3 3 dt R R 3q"n u t ! 3t # 3Mi , (C.5) p 3 3d ¸ 3q ¸ 4q dt 3 3 where i, u, t denote current, voltage and #ux linkage respectively. Subscripts r and s stand for rotor and stator. u is the rotor speed. d and q denote (`directa and 3 `quadraturea) components of the vectors with respect to the "xed stator reference frame (sD, sQ). ¸ and R are the autoinductances and resistances, M is the mutual inductance and p"1!(M2/¸ ¸ ). t is the load torque. 4 3 L C.1. Machine data used in simulations ASEA 3&50 Hz(ABB) MBL 132 SB38-2 7.5 kW. n number of pole pairs 1. p R stator resistance 2.19 ). 4 R rotor resistance 1.038 ). 3 ¸ stator autoinductance 0.51159 H. 4 ¸ rotor autoinductance 0.51159 H. 3 M mutual inductance 0.501 H. J rotor inertia 0.35 kg m2. G.J. Kulawski, M.A. Brdys& / Automatica 36 (2000) 5}22 References BrdysH , M. A., & Du, T. (1991). Algorithms for joint state and parameter estimation in induction motor drive systems. In Proceedings of the international conference CONTROL'91, Edinburgh, Scotland (pp. 915}920). BrdysH , M. A., Kulawski, G. J., & Quevedo, J. (1996). Recurrent networks for nonlinear adaptive control. In Proceedings of the 13th IFAC world congress, vol. F. San Francisco (pp. 151}156). BrdysH , M. A., Kulawski, G. J., & Quevedo, J. (1998). Recurrent networks for nonlinear adaptive control. IEE Proceedings on Control Theory and Applications, 145(2), 177}188. Chang, W. D., Fu, L. C., & Yang, J. H. (1996). Adaptive robust neural-network based control for siso systems. In Proceedings of the 13th IFAC world congress, vol. F. San Francisco (pp. 163}168). Chen, F. C., & Khalil, H. K. (1991). Adaptive control of nonlinear systems using neural networks * a dead-zone approach. In Proceedings of the American control conference (pp. 667} 672). Cybenko, G. (1989). Approximation by superpositions of a sigmoidal function. Mathematics of Control, Signals and Systems, 2, 303}314. Delgado, A., Kambhapati, C., & Warwick, K. (1995). Dynamic recurrent neural network for system identi"cation and control. IEE Proceedings on Control Theory and Applications, 142, 307}314. Du, T., & BrdysH , M. A. (1993). Implementations of extended Luenberger observers for joint state and parameter estimation of pwm induction motor drive. In Proceedings of the 5th European conference on power electronics (IPE), Brighton, England. Funahashi, K. I. (1989). On the approximate realization of continuous mappings by neural networks. Neural Networks, 2, 183}191. Funahashi, K. I., & Nakamura, Y. (1993). Approximation of dynamical systems by continuous time recurrent networks. Neural Networks, 6, 801}806. Gupta, M. M., Rao, D. H., & Nikiforuk, P. N. (1993). Dynamic neural network based inverse kinematic transformation of two- and threelinked robots. In Proceedings of the 12th IFAC world congress, vol. 3. Sydney (pp. 289}296). Hornik, K., Stinchcombe, M., & White, H. (1989). Multilayer feedforward networks are universal approximators. Neural Networks, 2, 359}366. Jin, L., Nikiforuk, P. N., & Gupta, M. M. (1993). Model matching control of unknown nonlinear systems using recurrent neural networks. In Proceedings of the 12th IFAC world congress, vol. 1. Sydney (pp. 337}344). Jin, L., Nikiforuk, P. N., & Gupta, M. M. (1994). Adaptive control of discrete-time nonlinear systems using recurrent neural networks. IEE Proceedings on Control Theory and Applications, 141, 169}176. Jin, L., Nikiforuk, P. N., & Gupta, M. M. (1995). Approximation of discrete-time state-space trajectories using dynamic recurrent neural networks. IEEE Transactions on Automatic Control, 40, 1266}1270. Khalil, H. K. (1992). Nonlinear systems. New York: Macmillan. KrsticH , M., Kanellakopoulos, I., & KokotovicH , P. (1995). Nonlinear and adaptive control design. New York: Wiley. Ku, C. C., & Lee, K. Y. (1992). System identi"cation and control using diagonal recurrent neural networks. In Proceedings of the American control conference (pp. 545}549). Kulawski, G. J., & BrdysH , M. A. (1994). Dynamic neural networks for nonlinear adaptive control. Technical Report 11/04-94. School of Electronic and Electrical Engineering, The University of Birmingham, UK. Levin, A. U., & Narendra, K. S. (1993). Control of nonlinear dynamical systems using neural networks: Controllability and stabilisation. IEEE Transactions on Neural Networks, 4, 192}206. 21 Levin, A. U., & Narendra, K. S. (1996). Control of nonlinear dynamical systems using neural networks * Part II: Observability, identi"cation and control. IEEE Transactions on Neural Networks, 7, 30}42. Lewis, F. L., Yesildirek, A., & Liu, K. (1993). Neural net robot controller: Structure and stability proofs. In Proceedings of the 32nd conference on decision and control, San Antonio (pp. 2785}2791). Marino, R., & Tomei, P. (1995). Nonlinear control design: Geometric, adaptive and robust. Englewood Cli!s, NJ: Prentice-Hall. Marino, R., Peresada, S., & Valigi, P. (1993). Adaptive input-output linearizing control of induction motors. IEEE Transactions on Automatic Control, 38(2), 208}221. Narendra, K. S., & Annaswamy, A. M. (1989). Stable adaptive systems. Englewood Cli!s, NJ: Prentice-Hall. Narendra, K. S., & Parthasarathy, K. (1990). Identi"cation and control of dynamical systems using neural networks. IEEE Transactions on Neural Networks, 1, 4}27. Narendra, K. S., & Parthasarathy, K. (1991). Gradient methods for the optimization of dynamical systems containing neural networks. IEEE Transactions on Neural Networks, 2, 252}262. Parlos, A. G., Chong, K. T., & Atiya, A. F. (1994). Application of the recurrent multilayer perceptron in modelling complex process dynamics. IEEE Transactions on Neural Networks, 5(2), 255}266. Sanner, R. M., & Slotine, &J. -J. E. (1992). Gaussian networks for direct adaptive control. IEEE Transactions on Neural Networks, 3(6), 837}863. Sastry, P. S., Santharam, G., & Unnikrishnan, K. P. (1994). Memory neuron networks for identi"cation and control of dynamical systems. IEEE Transactions on Neural Networks, 5(2), 306}319. Sastry, S., & Bodson, M. (1989). Adaptive control: Stability, convergence and robustness. Englewood Cli!s, NJ: Prentice-Hall. Suykens, J. A. K., Vandewalle, J., & De Moor, B. L. R. (1997). Nonlinear H control for continuous-time recurrent neural networks. In Pro= ceedings of the fourth European control conference, Brussels. Vas, P. (1990). Vector control of AC machines. Oxford: Oxford University Press. Verrelst, H., Van Acker, K., Suykens, J. A. K., Motmans, B., De Moor, B. L.R., & Vandewalle, J. (1997). N¸ neural control theory: Case q study for a ball and beam system. In Proceedings of the fourth European control conference, Brussels. Grzegorz Kulawski was born in Poland in 1970. He obtained a B.Eng. in Electronic and Electrical Engineering in 1993 and a PhD in 1998, both from The University of Birmingham, UK. At present he is with Shell International Exploration and Production B.V., Research and Technical Services, Rijswijk, The Netherlands. His research interests include neural networks, nonlinear adaptive control and modelling of industrial processes. Mietek Brdys: received the M.Sc. degree in Electronic Engineering and the Ph.D. and the D.Sc. degrees in Control Systems from the Institute of Automatic Control at the Warsaw University of Technology in 1970, 1974 and 1980, respectively. From 1974 to 1983, he held the posts of Assistant Professor and Associate Professor at the Warsaw University of Technology. In 1992 he became Full Professor of Control Systems in Poland. Between 1978 and 1995, he held 22 G.J. Kulawski, M.A. Brdys& / Automatica 36 (2000) 5}22 various visiting faculty positions at the University of Minnesota, City University, De Montfort University and University Polytecnic of Catalonia. Since January 1989, he has held the post of Senior Lecturer in the School of Electronic and Electrical Engineering at The University of Birmingham, UK. He has served as the Consultant for the Honeywell Systems and Research Center in Minneapolis, GEC Marconi and Water Authorities in UK, France, Germany, Spain and Poland. His research is supported by the UK Research Council and industry and the European Commission. He is the author or co-author of about 100 refereed paper and "ve books. His current research interests include intelligent control of nonlinear and uncertain systems, robust monitoring and operational control with application to environmental systems. He is a Chartered Engineer, a Member of the IEE and the IEEE, a Fellow of IMA and a member of IFAC Technical Committee on Large Scale Systems.
© Copyright 2025 Paperzz