• Nie Znaleziono Wyników

Optimal controller/observer gains of discounted-cost LQG systems

N/A
N/A
Protected

Academic year: 2021

Share "Optimal controller/observer gains of discounted-cost LQG systems"

Copied!
6
0
0

Pełen tekst

(1)

Delft University of Technology

Optimal controller/observer gains of discounted-cost LQG systems

Bijl, Hildo; Schön, Thomas B. DOI

10.1016/j.automatica.2018.12.040 Publication date

2019

Document Version

Accepted author manuscript Published in

Automatica

Citation (APA)

Bijl, H., & Schön, T. B. (2019). Optimal controller/observer gains of discounted-cost LQG systems. Automatica, 101, 471-474. https://doi.org/10.1016/j.automatica.2018.12.040

Important note

To cite this publication, please use the final published version (if applicable). Please check the document version above.

Copyright

Other than for strictly personal use, it is not permitted to download, forward or distribute the text or part of it, without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license such as Creative Commons. Takedown policy

Please contact us and provide details if you believe this document breaches copyrights. We will remove access to the work immediately and investigate your claim.

This work is downloaded from Delft University of Technology.

(2)

Optimal controller/observer gains of discounted-cost LQG

systems

Hildo Bijl

a

, Thomas B. Sch¨

on

b

aDelft Center for Systems and Control, Delft University of Technology, The Netherlands bDepartment of Information Technology, Uppsala University, Sweden

Abstract

The linear-quadratic-Gaussian (LQG) control paradigm is well-known in literature. The strategy of minimizing the cost function is available, both for the case where the state is known and where it is estimated through an observer. The situation is different when the cost function has an exponential discount factor, also known as a prescribed degree of stability. In this case, the optimal control strategy is only available when the state is known. This paper builds on from that result, deriving an optimal control strategy when working with an estimated state. Expressions for the resulting optimal expected cost are also given.

Key words: Linear systems, cost function, LQG, optimal control, Riccati equation.

1 Introduction

Consider the continuous-time linear system1

˙

x(t) = Ax(t) + Bu(t) + v(t), (1a) y(t) = Cx(t) + Du(t) + w(t), (1b) with x the state, u the input, y the output, v and w Gaussian white noise with respective intensities V and W , and A, B, C and D the system matrices. We assume that the initial state x0 is unknown but

dis-tributed according to a Gaussian with µ0= E[x0] and

Σ0= Ex0x0T. Note that Σ0is not the variance of x0.

Our goal is to control system (1) such as to minimize the discounted (exponential) quadratic cost function

J (T ) = E " Z T 0 e2αt xT(t)Qx(t) + uT(t)Ru(t) dt # , (2)

Email addresses: h.j.bijl@tudelft.nl (Hildo Bijl), thomas.schon@it.uu.se (Thomas B. Sch¨on).

1 From a formal point of view the system notation of (1)

is incorrect, because v(t) and w(t) are not measurable with nonzero probability. However, since this notation is common in the control literature, we will stick with it. For methods to properly deal with stochastic differential equations, see [11].

with J (T ) the expected cost, the real number α the discount exponent/prescribed degree of stability, and Q ≥ 0 and R > 0 symmetric weight matrices. In partic-ular, we will optimize the infinite-time expected cost J , with

J = lim

T →∞J (T ). (3)

Our contribution in this paper is that we derive the op-timal controller and observer gains for the continuous-time linear system (1) such that the expected cost J given in (3) is minimized.

2 Related work

Linear-Quadratic-Gaussian (LQG) systems—linear sys-tems with a quadratic cost function subject to Gaussian noise—have been thoroughly investigated in the past. This was especially true near the 1960s, with for instance the publication of the Kalman filter [8,7].

The discoveries from the decades afterwards have been summarized in numerous textbooks. Examples include the books by [12, Chapter 7], [9, Chapter 5], [6, Chapter 1], [2, Chapters 3, 8], [15, Chapter 6], [17, Chapter 10], [14, Chapter 9] and [4, Chapter 4]. All these books exam-ine the non-discounted cost function (with α = 0), save for [2, Section 3.5] that also considers the discounted cost function, presenting results from an earlier paper [1]. Here it was shown that discounting the cost function is equivalent to prescribing a degree of stability.

(3)

The prescribed degree of stability is actually a relevant problem in that it is a generalization of the regular LQG paradigm with the non-discounted cost function. There is also a variety of applications of this idea, such as fault tolerant flight control [5], spacecraft guidance [10] and robot manipulators [16]. However, to the best of the authors’ knowledge there are still fundamental proper-ties remaining to be established and our contribution in this paper is to provide one of those. The work [1] ex-amined the situation where the state is assumed to be known. If the state can only be observed through noisy output measurements—a familiar problem for the non-discounted cost function—then we are not aware of any work that jointly optimize the controller and the state estimator. The closest is the work by [13], who strived to derive a state estimator with minimal mean squared er-ror, given a prescribed convergence rate. However, that work ignored the uncertainty in the initial state and did not examine the problem of jointly optimizing the con-troller and observer gains. In fact, it was not mentioned whether the separation principle still holds or not when using the discounted cost function. Hence, the problem of jointly optimizing the controller and observer gains, subject to a discounted cost function and an uncertain initial state, appears to be an open problem.

3 Brief summary of known theorems

To place our new result in perspective, we briefly ex-amine some known results first. We start with the non-discounted case (α = 0) where the state x(t) is known (i.e., C = I and W = 0). In this case the optimal control law is given by the following theorem.

Theorem 1 Consider system (1), where the state is as-sumed known. If (A, B) is stabilizable, then the optimal control law minimizing the expected non-discounted cost J (i.e., with α = 0) is a linear control law u(t) = −F x(t), where

F = R−1BTX, (4) and X is the solution to the Riccati equation

ATX + XA + Q − XBR−1BTX = 0. (5) When V = 0, the corresponding expected cost equals

J = Ex0TXx0 = tr (XΣ0) . (6)

When V 6= 0, then J (T ) → ∞, but the steady-state cost rate equals

lim

T →∞

dJ (T )

dT = tr (XV ) . (7)

PROOF. See any of the aforementioned books; for ex-ample [9, Theorem 3.9].

There is another way to look at the Theorem 1, which will become important in the proof of our main result. We know from [3, Theorem 3] that, for the above situation, and for any feedback matrix F , the expected steady-state cost rate equals

lim

T →∞

dJ (T )

dT = tr (XV ) , (8) where X per definition is the unique solution to the Lya-punov equation

(A − BF )TX + X(A − BF ) + Q + FTRF = 0. (9) To minimize the above cost rate, we must find the value of F minimizing (8). Theorem 1 tells us that the cost rate (8) is minimized when X satisfies (5) and F sub-sequently equals (4). This is irrespective of the value of the positive definite matrix V .

Next, consider the case where there is a discount expo-nent α 6= 0. Now the solution is given by the following Theorem. Note that α can be positive (a prescribed de-gree of stability) or negative (a discount exponent), but for ease of writing we always call it a discount exponent. Theorem 2 Consider system (1), where the state is as-sumed known. Define Aα= A+αI. If (Aα, B) is

stabiliz-able, then the optimal control law minimizing the expected discounted cost J is a linear control law u(t) = −Fαx(t),

where

Fα= R−1BTXα, (10)

and Xαis the solution to the Riccati equation

ATαXα+ XαAα+ Q − XαBR−1BTXα= 0. (11)

The corresponding expected cost (for both zero and nonzero V ) when α < 0 equals

J = tr  Xα  Σ0− V 2α  . (12) When α ≥ 0, then J (T ) → ∞ as T → ∞.

PROOF. A proof is given by [2, Section 3.5].

When the state is unknown, an observer needs to be used. The state estimate ˆx of this observer is updated through

˙ ˆ

x(t) = Aˆx(t)+Bu(t)+K (y(t) − C ˆx(t) − Du(t)) , (13) subject to some initial state estimate ˆx(0). If the state estimation error e(t) is defined as e(t) = ˆx(t)−x(t), then this error (i.e., its variance) can be minimized through the following Theorem.

(4)

Theorem 3 Consider system (1). If (A, C) is de-tectable, then the optimal observer gain minimizing the steady-state error covariance is

K = ECTW−1, (14) where E is the optimal steady-state error covariance, found through

AE + EAT + V − ECTW−1CE = 0. (15)

PROOF. This is the famous Kalman-Bucy filter from [8]. A proof can also be found in [9, Theorem 4.5].

The above result holds regardless of the value of α, be-cause it is unrelated to the cost J . If our goal is to op-timize the cost J subject to α = 0 (the non-discounted case) then the following Theorem provides the solution. Theorem 4 Consider system (1). If (A, B) is stabiliz-able and (A, C) is detectstabiliz-able, then the optimal control law minimizing the expected non-discounted cost (i.e., with α = 0) is a linear control law u(t) = −F ˆx(t), with F given by (4), ˆx(t) following from (13) and the observer gain K taken as (14). The resulting expected steady-state cost rate is given by

lim T →∞ dJ (T ) dT = tr XKW K T + EQ = tr XV + EFTRF , (16) with X the solution of (5) and E the solution of (15).

PROOF. The optimal controller and observer gains follow from the separation principle. See for instance [9, Theorem 5.4]. Expressions for the expected steady-state cost rate can be derived using [3, Theorem 3].

4 Optimizing the discounted cost function In this section we derive the main result: the optimal controller/observer gains minimizing the discounted cost function, subject to an unknown state. It is important to realize that ‘optimal’ here only means that the expected discounted cost (2) is minimized. There is no guarantee that the steady-state error variance, or any other quan-tity, is still at a minimum.

Theorem 5 Consider system (1). If (Aα, B) is

stabiliz-able and (Aα, C) is detectable, then the optimal control

law minimizing the expected discounted cost J is a linear control law u(t) = −Fαx(t), with Fˆ αgiven by (10) and

Xαgiven by (11). Identically to (13), ˆx(t) is provided by

the observer ˙

ˆ

x(t) = Aˆx(t)+Bu(t)+Kα(y(t) − C ˆx(t) − Du(t)) , (17)

where ˆx0is set to µ0, the observer gain Kαis given by

Kα= EαCTW−1 (18)

and Eαis the solution to the Riccati equation

AαEα+ EαAαT + V − 2α Σ0− µ0µ0T

−EαCTW−1CEα= 0. (19)

The corresponding expected cost for α < 0 equals

J = 1 −2αtr XαKαW K T α + EαQ + µ0TXαµ0 = 1 −2αtr XαV + EαF T αRFα + tr (XαΣ0) . (20) When α ≥ 0, then J (T ) → ∞ as T → ∞.

PROOF. To start, we write the joint dynamics of the system and its observer as

" ˙ x(t) ˙ e(t) # = " A − BFα −BFα 0 A − KαC # " x(t) e(t) # + " v(t) Kαw(t) − v(t) # = ˜A˜x(t) + ˜v(t), (21) and the total expected cost as

J (T ) = E " Z T 0 e2αtx˜T(t) ˜Q˜x(t) dt # . (22)

Note that the tilde-notation used above denotes proper-ties of the joint dynamics. We have already defined ˜A, ˜

x(t) and ˜v(t) as above. The variance ˜V of ˜v, the mean and variance of ˆx0and the weight matrix ˜Q satisfy

˜ V = " V −V −V KαW KαT + V # , (23a) ˜ µ0= E [˜x0] = E " x0 e0 # = " µ0 0 # , (23b) ˜ Σ0= E h ˜ x0x˜0T i = " µ0µ0T µ0µ0T − Σ0 µ0µ0T − Σ0 Σ0− µ0µ0T # , (23c) ˜ Q = " Q + FT αRFα FαTRFα FαTRFα FαTRFα # . (23d)

(5)

Our goal is to choose Fαand Kαso as to minimize the

expected cost J . This cost, according to [3, Theorem 2], equals J = tr X˜α Σ˜0− ˜ V 2α !! , (24)

where ˜Xαper definition is the unique solution to

˜

ATαX˜α+ ˜XαA˜α+ ˜Q = 0, (25)

and where ˜Aαis defined as ˜A+αI. Expression (24) holds

for any Fαand Kα, which implies that we need to find

the Fα and Kα that minimize it. Note that we cannot

directly solve this by applying Theorem 1, because this time ˜V is not constant: it depends on Kα. We need a

different method.

First, we expand our matrix equations into elements. This turns (25) into the following three equations,

(Aα− BFα)TX˜α11+ ˜X 11 α (Aα− BFα) + Q + FαTRFα = 0, (26a) (Aα− BFα)TX˜α12− ˜X 11 α BFα + ˜Xα12(Aα− KαC) + FαTRFα= 0, (26b) (Aα− KαC) T ˜ Xα22− FαTB TX˜12 α − ˜Xα21BFα+ ˜Xα22(Aα− KαC) + FαTRFα= 0. (26c)

There is a fourth expression, but it is identical to (26b) — to be precise, it is its transpose — so it is not worth mentioning. Similarly we can expand (24) as

J = tr  ˜ Xα11  Σ0− V 2α  − ˜Xα12  Σ0− µ0µ0T− V 2α  − ˜Xα21  Σ0− µ0µ0T − V 2α  + ˜Xα22  Σ0− µ0µ0T − V 2α − KαW KαT 2α  . (27)

It is difficult to jointly optimize Fαand Kαto minimize

the above cost. The key here is to first assume a certain value for Fαand then find the value of Kαthat is optimal

for this particular value of Fα. To be precise, we assume

that Fαis given by (10).

It is interesting to note that this value for Fα happens

to optimize the first term of (27),

J11= tr  ˜ Xα11  Σ0− V 2α  . (28)

After all, from (26a) we see that ˜Xα11 solely depends

on Fα and not on Kα. The problem of optimizing Fα

now turns out to be equivalent to the problem solved by Theorem 2. (Also see the note after Theorem 1.) It

follows that the value of Fαminimizing J11equals (10),

and that ˜Xα11from (26a) equals the solution Xαof (11).

For this assumed value of Fα, the other equations greatly

simplify. If we insert (10) into (26b), we directly find that ˜X12

α = 0. This tells us that the separation principle

still holds for this situation, albeit in an adjusted form. At the same time (26c) reduces to

(Aα− KαC)TX˜α22+ ˜X 22

α (Aα− KαC) + FαTRFα= 0.

(29) Our goal is to find the value of Kαminimizing the last

term from (27). That is, we want to minimize

J22= tr  ˜ Xα22  Σ0− µ0µ0T − V 2α− KαW KαT 2α  . (30) According to [3, Theorem 16], we can rewrite this as

J22= tr  −Eα 2α  FαTRFα  , (31)

where the term −Eα

2α per definition satisfies

(Aα− KαC)  −Eα 2α  +  −Eα 2α  (Aα− KαC) T +  Σ0− µ0µ0T − V 2α  −KαW K T α 2α = 0. (32) From this, we can directly find the value of Kα

mini-mizing (31), and hence minimini-mizing the total expected cost J . According to the principle described right after Theorem 1, it equals (18).

To summarize, we have assumed that Fα was given

by (10) and subsequently found that the optimal value for Kα equals (18). Of course this does not

necessar-ily mean that this combination of Fα and Kα jointly

optimizes the expected cost J . We need one more step. For this step, we have to reverse the process: first we as-sume that Kαequals (18) and subsequently we optimize

the cost J for Fα. When doing so, we do have to use

a different system notation. Instead of considering the joint dynamics of x and e in (21), we consider the joint dynamics of ˆx and e. Additionally, instead of optimizing the cost J written as (24), we first use [3, Theorem 16] to rewrite the expression. If we follow these steps, then in an identical way we find that the optimal value of Fα

equals (10).

To conclude, if we choose Fαas (10) then the optimal Kα

equals (18), and vice versa if we choose Kαas (18) then

the optimal Fαequals (10). This proves that our

combi-nation of (Fα, Kα) is at least a local solution to the

op-timization problem. However, because the opop-timization

(6)

problem is convex in both Fα and Kα, it must also be

the global optimum. Hence, we conclude that the com-bination of Fαand Kαminimizes the expected cost J .

The only thing left to prove is the cost expression (20). The second line from this equation follows directly from (28) and (31): just add J11 and J22. The first line

follows in the same way, if you redo the full derivation with the joint state of ˆx and e, as described above. That concludes this proof.

This theorem shows how to optimally trade off between compensating for process noise (V ), for measurement noise (W ) and for uncertainty in the initial state (Σ0−

µ0µ0T). None of the previously derived theorems had

to include all these three parameters in their trade-off, which is what makes this new result significant.

Due to the separation principle, the stability of the con-trolled system is similar to when we applied Theorem 2. The eigenvalues of the closed-loop system are all guar-anteed to be smaller than −α [2, Section 3.5]. Hence, if α > 0, stability is guaranteed.

5 Conclusions and recommendations

Through Theorem 5 it is now possible to find the op-timal controller and observer gains of an LQG system with discounted cost. This paper also serves as a focused overview of this part of control engineering.

Future work on this subject can look into replacing the discount exponent α by a discount matrix, investigate the effect of a finite time window T on the optimal con-troller/observer parameters, or examine time-varying systems, similarly to [13], to see whether the same results are still applicable.

Acknowledgements

This research is supported by the Dutch Technology Foundation STW, which is part of the Netherlands Or-ganisation for Scientific Research (NWO), and which is partly funded by the Ministry of Economic Affairs (Project number: 12173, SMART-WIND). The work was also supported by the Swedish research Council (VR) via the project NewLEADS - New Directions in Learning Dynamical Systems (Contract number: 621-2016-06079) and by the Swedish Foundation for Strategic Research (SSF) via the project ASSEMBLE (Contract number: RIT15-0012).

References

[1] Brian D. O. Anderson and John B. Moore. Linear system optimisation with prescribed degree of stability. In

Proceedings of the Institution of Electrical Engineers, volume 116, pages 2083–2087, 1969.

[2] Brian D. O. Anderson and John B. Moore. Optimal Control: Linear Quadratic Methods. Prentice Hall, 1990.

[3] Hildo Bijl, Jan-Willem van Wingerden, Thomas B. Sch¨on, and Michel Verhaegen. Mean and variance of the LQG cost function. Automatica, 67:216–223, May 2016.

[4] Okko H. Bosgra, Huibert Kwakernaak, and Gjerrit Meinsma. Design Methods for Control Systems. Dutch Institute of Systems and Control (DISC), 2008.

[5] Bogdan D. Ciubotaru, Marcel Staroswiecki, and Nicolai D. Christov. Extended hybrid technique for control redesign with stabilization and correction. In American Control Conference (ACC), pages 5152–5158, June 2013.

[6] Michael J. Grimble and Michael A. Johnson. Optimal Control and Stochastic Estimation. Wiley, 1988.

[7] Rudolf E. Kalman. A new approach to linear filtering and prediction problems. Journal of Basic Engineering, 82:35– 45, 1960.

[8] Rudolf E. Kalman and Richard S. Bucy. New results in linear filtering and prediction theory. Journal of Basic Engineering, 83:95–107, 1961.

[9] Huibert Kwakernaak and Raphael Sivan. Linear optimal control systems. Wiley Interscience, 1972.

[10] Yunhe Meng, Qifeng Chen, and Qing Ni. A new geometric guidance approach to spacecraft near-distance rendezvous problem. Acta Astronautica, 129:374–383, 2016.

[11] Bernt Øksendal. Stochastic Differential Equations. Springer-Verlag, 1985.

[12] Karl Johan ˚Astr¨om. Introduction to Stochastic Control Theory. Academic Press, 1970.

[13] Ilan Rusnak. Least mean squares error based filter of linear system with prescribed convergence rate. In IEEE International Conference on the Science of Electrical Engineering (ICSEE), pages 1–5, August 2016.

[14] Sigurd Skogestad and Ian Postlethwaite. Multivariable Feedback Control: Analysis and Design. John Wiley & Sons, 2005.

[15] Robert F. Stengel. Optimal Control and Estimation. Dover Publications, 1994.

[16] Nobuya Takahashi and Osamu Sato. Guaranteed cost control of robot manipulator with prescribed degree of stability. Artificial Life and Robotics, 21(4):520–524, August 2016. [17] Harry L. Trentelman, Anton A. Stoorvogel, and Malo Hautus.

Cytaty

Powiązane dokumenty

Assuming symmetric properties of the observer after rotation direction change, vectors of real and estimated stator current and rotor flux for negative speed values (Figure 7b)

It consists of a closed-loop observer, an integral controller, the object, state feedback and estimated controlled outputs feedback [7]..

Let X be a Banach space and let k\(X) denote the inflmum of all numbers k such that there exists a retraction of the unit ball onto the unit sphere being a ¿-set

3) przyszłą liczbę urodzeń (na podstawie przewidywania przyszłych współczynników płodności), 4) wynikowe współczynniki demograficzne. W niektórych przypadkach może

d) prawa i obowiązki podmiotów sektora informacyjnego wobec obywateli, pań- stwa i pozostałych podmiotów społecznych i gospodarczych (Oleński, 2006). wypełnianie przez obywateli

When compared with the relevant past works on optimal control of an aircraft, the proposed controller does not depend on the aircraft dynamics via the design of an

This paper will briefly outline some of the characteristics of the social changes that have occurred over the past fifty years and which have had a significant impact on the

To, co jest tu istotne spro- wadza się do następującej tezy: niezależnie od odmienności termino- logii, języków i tradycji badawczej, prace Janusza Sławińskiego pisane