)_, [s -I J-' I [-,+a IJ [~ s(!+a)]

Similar documents
ECEN 420 LINEAR CONTROL SYSTEMS. Lecture 6 Mathematical Representation of Physical Systems II 1/67

THE REACTION WHEEL PENDULUM

Motor Controller. A block diagram for the motor with a feedback controller is shown below

10/8/2015. Control Design. Pole-placement by state-space methods. Process to be controlled. State controller

MATRIX ALGEBRA AND SYSTEMS OF EQUATIONS. + + x 1 x 2. x n 8 (4) 3 4 2

Quanser NI-ELVIS Trainer (QNET) Series: QNET Experiment #02: DC Motor Position Control. DC Motor Control Trainer (DCMCT) Student Manual

Appendix A: Exercise Problems on Classical Feedback Control Theory (Chaps. 1 and 2)

Time Response Analysis (Part II)

Contents. PART I METHODS AND CONCEPTS 2. Transfer Function Approach Frequency Domain Representations... 42

Represent this system in terms of a block diagram consisting only of. g From Newton s law: 2 : θ sin θ 9 θ ` T

Determinants Chapter 3 of Lay

(Refer Slide Time: 00:01:30 min)

Chapter 3. State Feedback - Pole Placement. Motivation

The output voltage is given by,

Introduction to Feedback Control

1 Matrices and matrix algebra

Control Systems. Design of State Feedback Control.

Advanced Control Theory

Dr Ian R. Manchester Dr Ian R. Manchester AMME 3500 : Review

Laboratory 11 Control Systems Laboratory ECE3557. State Feedback Controller for Position Control of a Flexible Joint

FUZZY LOGIC CONTROL Vs. CONVENTIONAL PID CONTROL OF AN INVERTED PENDULUM ROBOT

R a) Compare open loop and closed loop control systems. b) Clearly bring out, from basics, Force-current and Force-Voltage analogies.

School of Mechanical Engineering Purdue University. ME375 Feedback Control - 1

Topic 15 Notes Jeremy Orloff

(b) A unity feedback system is characterized by the transfer function. Design a suitable compensator to meet the following specifications:

Video 5.1 Vijay Kumar and Ani Hsieh

C(s) R(s) 1 C(s) C(s) C(s) = s - T. Ts + 1 = 1 s - 1. s + (1 T) Taking the inverse Laplace transform of Equation (5 2), we obtain

Linear Second Order ODEs

MULTIVARIABLE ZEROS OF STATE-SPACE SYSTEMS

EEE 184: Introduction to feedback systems

Positioning Servo Design Example

[Disclaimer: This is not a complete list of everything you need to know, just some of the topics that gave people difficulty.]

Test 2 SOLUTIONS. ENGI 5821: Control Systems I. March 15, 2010

Table of Laplacetransform

The Laplace Transform

Control Systems Design


Linear State Feedback Controller Design

Control of Electromechanical Systems

Lecture 2: Discrete-time Linear Quadratic Optimal Control

Module 07 Controllability and Controller Design of Dynamical LTI Systems

Lab 6a: Pole Placement for the Inverted Pendulum

Things we can already do with matrices. Unit II - Matrix arithmetic. Defining the matrix product. Things that fail in matrix arithmetic

Control Systems I. Lecture 7: Feedback and the Root Locus method. Readings: Guzzella 9.1-3, Emilio Frazzoli

Laboratory Exercise 1 DC servo

Digital Control: Part 2. ENGI 7825: Control Systems II Andrew Vardy

What is A + B? What is A B? What is AB? What is BA? What is A 2? and B = QUESTION 2. What is the reduced row echelon matrix of A =

EE C128 / ME C134 Fall 2014 HW 9 Solutions. HW 9 Solutions. 10(s + 3) s(s + 2)(s + 5) G(s) =

Problem Set 3: Solution Due on Mon. 7 th Oct. in class. Fall 2013

FEEDBACK CONTROL SYSTEMS

Linear Algebra: Lecture Notes. Dr Rachel Quinlan School of Mathematics, Statistics and Applied Mathematics NUI Galway

EE 422G - Signals and Systems Laboratory

Control Systems I. Lecture 6: Poles and Zeros. Readings: Emilio Frazzoli. Institute for Dynamic Systems and Control D-MAVT ETH Zürich

KINGS COLLEGE OF ENGINEERING DEPARTMENT OF ELECTRONICS AND COMMUNICATION ENGINEERING

Outline. Classical Control. Lecture 2

Kevin James. MTHSC 3110 Section 2.1 Matrix Operations

Renormalization Group analysis of 2D Ising model

Proportional plus Integral (PI) Controller

CONTROL DESIGN FOR SET POINT TRACKING

S I X SOLUTIONS TO CASE STUDIES CHALLENGES. Antenna Control: Stability Design via Gain K s s s+76.39K. T(s) =

Modeling and System Identification for a DC Servo

(a) Torsional spring-mass system. (b) Spring element.

7.2 Relationship between Z Transforms and Laplace Transforms

Here represents the impulse (or delta) function. is an diagonal matrix of intensities, and is an diagonal matrix of intensities.

Matrix Theory and Differential Equations Homework 6 Solutions, 10/5/6

MAS107 Control Theory Exam Solutions 2008

Time Response of Systems

Appendix W. Dynamic Models. W.2 4 Complex Mechanical Systems. Translational and Rotational Systems W.2.1

MTH Linear Algebra. Study Guide. Dr. Tony Yee Department of Mathematics and Information Technology The Hong Kong Institute of Education

PID Control. Objectives

Control Systems. Laplace domain analysis

conventions and notation

Quantitative Understanding in Biology Module III: Linear Difference Equations Lecture II: Complex Eigenvalues

Control Systems I. Lecture 7: Feedback and the Root Locus method. Readings: Jacopo Tani. Institute for Dynamic Systems and Control D-MAVT ETH Zürich

Example: 2x y + 3z = 1 5y 6z = 0 x + 4z = 7. Definition: Elementary Row Operations. Example: Type I swap rows 1 and 3

Control of Manufacturing Processes

1 GSW Sets of Systems

Matrix inversion and linear equations

MAT 1332: CALCULUS FOR LIFE SCIENCES. Contents. 1. Review: Linear Algebra II Vectors and matrices Definition. 1.2.

EECS C128/ ME C134 Final Wed. Dec. 15, am. Closed book. Two pages of formula sheets. No calculators.

EE Experiment 11 The Laplace Transform and Control System Characteristics

Chapter 9 Observers, Model-based Controllers 9. Introduction In here we deal with the general case where only a subset of the states, or linear combin

ECE-320: Linear Control Systems Homework 8. 1) For one of the rectilinear systems in lab, I found the following state variable representations:

EE C128 / ME C134 Final Exam Fall 2014

ME 132, Fall 2017, UC Berkeley, A. Packard 334 # 6 # 7 # 13 # 15 # 14

Select/Special Topics in Atomic Physics Prof. P. C. Deshmukh Department of Physics Indian Institute of Technology, Madras

Question: Total. Points:

Lecture 12. Upcoming labs: Final Exam on 12/21/2015 (Monday)10:30-12:30

EC 8391-CONTROL SYSTEMS ENGINEERING. Questions and Answers PART-A. Unit - I Systems Components And Their Representation

Root Locus. Motivation Sketching Root Locus Examples. School of Mechanical Engineering Purdue University. ME375 Root Locus - 1

Bangladesh University of Engineering and Technology. EEE 402: Control System I Laboratory

Mathematical Modeling and Dynamic Simulation of a Class of Drive Systems with Permanent Magnet Synchronous Motors

Lec 6: State Feedback, Controllability, Integral Action

Stability of Feedback Control Systems: Absolute and Relative

Chapter 5 HW Solution

Stepping Motors. Chapter 11 L E L F L D

SRV02-Series Rotary Experiment # 1. Position Control. Student Handout

Feedback Control of Linear SISO systems. Process Dynamics and Control

ACCURATE FREE VIBRATION ANALYSIS OF POINT SUPPORTED MINDLIN PLATES BY THE SUPERPOSITION METHOD

Chapter Stability Robustness Introduction Last chapter showed how the Nyquist stability criterion provides conditions for the stability robustness of

Transcription:

70 CONTROL SYSTEM DESIGN 11! is known in mathematical literature as the resolvent of A. In engineering literature this matrix has been called the characteristic frequency matrix[ I] or simply the characteristic matrix.[ 4] Regrettably there doesn't appear to be a standard symbol for the resolvent, which we have designated as <I>(s) in this book. The fact that the state transition matrix is the inverse Laplace transform of the resolvent matrix facilitates the calculation of the former. It also characterizes the dynamic behavior of the system, the subject of the next chapter. The steps one takes in calculating the state-transition matrix using the resolvent are: (a) Calculate sf - A. (b) Obtain the resolvent by inverting (sf - A). (c) Obtain the state-transition matrix by taking the inverse Laplace transform of the resolvent, element by element. The following examples illustrate the process. Example 3C DC motor with inertial load In Chap. 2 (Example 2B) we found that the dynamics of a dc motor driving a n inertial load are O=w w = -aw + (3u The matrices of the state-space characterization are Thus the resolvent is )_, [s -I J-' I [-,+a IJ [~ s(!+a)] ClJ(s) = (sf - A = = --- = o s+a 5(s +a ) 0 s I o - s + a Finally, taking the inverse Laplace transforms of each term in ClJ(s) we obtain e A ' = <P(l) = [~ (1 - ee_-:,')/ a J Example 3D Inverted pendulum The equations of motion of an inverted pendulum were determined to be (approximately) O=w w=o'e +u Hence the matrices of the state-space characterization are B = [~J The resolvent is [ s -IJ-' I [s IJ _0 2 S s2-0 2 0 2 S ClJ(s) = (sf - A)-' = = - -

DYNAMICS 01"' L1N~AR SYSTEMS 71 and the state-transition matrix is A [COSh HI sinh lll/fl] <1>(1) = e '= H sinh HI cosh HI For a general kth-order system the matrix sf - A has the following appearance s - all - a 12 - alk J -a s - a... - Q k (3.50) sf - A =... 2.1.... 2~........ [ - akl - a u.. S - (/kk We recall (see Appendix) that the inverse of any matrix M can be written as the adjoint matrix, adj M, divided by the determinant IMI. Thus _,)- 1 adj (sf - A) ( s1 - A = -;:...;.--~ 1sT - AI If we imagine calculating the determinant Isf - AI we see that one of the terms will be the product of the diagonal elements of sf - A: (s - all)(s - an)'.. (s - akk) = Sk + CIS k - 1 +... -I- Ck a polynomial of degree k with the leading coefficient of unity. There will also be other terms coming from the off-diagonal elements of sf - A but none will have a degree as high as k. Thus we conclude that (3.51 ) This is known as the characteristic polynomial of the matrix A. It plays a vital role in the dynamic behavior of the system. The roots of this polynomial are called the characteristic roots, or the eigenvalues, or the poles, of the system and determine the essential features of the unforced dynamic behavior of the system, since they determine the inverse Laplace transform of the resolvent, which is the transition matrix. See Chap. 4. The adjoint of a k by k matrix is itself a k by k matrix whose elements are the cofactors of the original matrix. Each cofactor is obtained by computing the determinant of the matrix that remains when a row and a column of the original matrix are deleted. It thus follows that each element in adj (sf - A) is a polynomial in s of maximum degree k - 1. (The polynomial cannot have degree k when any row and column of sf - A is deleted.) Thus it is seen that the adjoint of sf - A can be written Thus we can express the resolvent in the following form - I Els k - I +... + Ek (sf - A) = -k;-' - - ;-k---:"i--- S + als +... + ak (3.52)

200 CONTROL SYSTEM DESIGN * >---- --+--y Figure 5.6 Block diagram showing that ba lanced bridge is neither controllable nor observable. (Elements with * open when bridge is balanced.) When numerical values are inserted for the physical parameters in the systems of Examples 5B and 5C there is no way of distinguishing between the qualitative nature of the uncontrollability of the two systems: they are both simply uncontrollable. But physically there is a very important distinction between the two systems. The two-mass mechanical system is uncontrollable for every value of the parameters (masses, spring rates); the only way to control the position of the center of mass is to add an external force. This necessitates a structural change to the system. The balanced bridge, however, is uncontrollable only for one specific relationship between the parameters, namely the balance condition (5CA). In other words, the system is almost always controllable. (As a practical matter, it will be difficult to control VI and V2 independently when (5CA) is nearly true. This raises the issue of degree of controllability, a topic discussed in Note 5.3.) It is important for the control system engineer to recognize this distinction, particularly when dealing with an unfamiliar process for which the state-space representation is given only by numerical data. A numerical error in calculating the elements of the A and B matrices, or an experimental error in measuring them, may make an uncontrollable system seem controllable. A control system designed with this data may seem to behave satisfactorily in simulation studies based on the erroneous design data, but will fail in practice. On the other hand, a process that appears to be uncontrollable (or nearly uncontrollable), but which is not structurally uncontrollable, may be rendered more tractable by changing some parameter of the process-by "unbalancing the bridge." Example 5D How not to control an unstable system (inverted pendulum) There are many ways of designing perfectly fine control systems for unstable processes such as the inverted pendulum of Examples 2E and 3D. These will be discussed at various places later on in this

CONTROLLABILITY AND OBSERVABILlTY 201 COMPENSATOR INVERTED PENDULUM u 8 Figure 5.7 Unstabilizable compensation of inverted pendulum. book. But one w~y guaranteed to be disastrous is to try to cancel the unstable pole with a zero in the compensator. The reason for the disaster is the subject of this example. Consider the inverted pendulum of Example 30 with the output being the measured position. The transfer function from the input (force) to the output (position) is yes) I 1 H(s) = -= --~=----- f(s) 52-0 2 (.I + O)(s - 0) (SO.I) This is obviously unstable. A much better transfer function would be I H(s)=-- sis + 0) (50.2) which is stable and, because of the pole at the origin, would be a "type-one" system, with zero steady state error, Thus, one might be tempted to.. compensate" the unstable transfer function by mea ns of a compensator having th e transfer function (Fig. 5.7) with s -.!1 0 G(s) = ---= 1- - s s (SO.3) Of course it will not be possible to make fl precisely equal to 0 so the compensation will not be perfect. But that is not the trouble, as we shall see. The compensator transfer function (50.3) represents" proportional plus integral" compensation which is quite customary in practical process control systems. The transfer function of the compensated system is now s -0 He(s) = G(s)H(s) = 2 2 -> H(s) sis - 0 ) as (SO.4) A block diagram representation of this system is shown in Fig. 5.7, and the state-space equations corresponding to this representation are X I = X 2 X 2 = 02XI - X; = flu X J + U (SO.5) where XJ is the state of the integrator in the compensator. The matrices of the process (50.5) are

202 CONTROL SYSTEM DESIGN n-o 2n2 I- --I@ u y Figure 5.8 Partial fraction representation of Fig. 5.7. The A matrix can be transformed c to diagonal form by the transformation matrix 0 I T=-I- 0 '] T-' - [~ i,j 2-0 -I -0 20 2 0 0 2 0 We find that 0 A - TAT' - [: -0 0 :] _ I _ [ n-u ] B = TB = -2 -(0 + 0) 20 _ 20 The state-space representation of the transformed system is as shown in Fig. 5.8. This block-diagram corresponds directly to the partial-fraction expansion of (50.4): [i/h 2 (0 - [i)/20 2 -(0 + [i)/2n 2 H (s) = --+ + --'---'---- c s s-o s+o (50.6) Note carefully what happens when n -> O. In the block-diagram the connection between the control input u 'and the unstable state x, is broken, rendering the system uncontrollable and unstabilizable. In (50.6) the residue at the unstable pole vanishes. But now we understand that the vanishing of a residue at a pole of a transfer function does not imply that the subsystem giving rise to the pole disappears, but rather that it becomes" invisible." If the original inverted pendulum could have arbitrary initial conditions, the transformed system (50.5) could also have arbitrary initial conditions and hence the inverted pendulum would most assuredly not remain upright, regardless of how the loop were closed between the measurement y and the control input u. More reasons for unobseevability The foregoing examples were instances of uncontrollable systems. Instances of unobservable systems are even more abundant. An unobservable system results any time a state variable is not measured

r CONTROLLABILITY AND OBSERVABILITY 203 -J f-----i. Y Figure 5.9 Systems in tandem that are unobservable. directly and is not fed back to those state variables that are measured. Thus, any system comprising two subsystems in tandem (as shown in Fig. 5.9, in which none of the states of the right-hand subsystem can be measured) is unobservable. The transfer function from the inputs to the outputs obviously depends only on the left-hand subsystem. Physical processes which have the structure shown in Fig. 5.9 are not uncommon. A mass m acted upon by a control force f is unobservable if only its velocity, and not its position, can be measured. This means that no method of velocity feedback can serve as a means of controlling position. In this regard it is noted that the integral of the measured velocity is not the same as the actual position. A control system shown in Fig. 5.10 will not be effective in controlling the position x of the mass, no matter how well it controls the velocity x; any initial position error will remain in the system indefinitely. In addition to the obvious reasons for unobservability there are also some of the more subtle reasons such as symmetry, as was illustrated by Example 5C. 5.3 DEFINITIONS AND CONDITIONS FOR CONTROLLABILITY AND OBSERV ABILITY In Secs. 5.1 and 5.2 we found that uncontrollable and! or unobservable systems were characterized by the property that the transfer function from the input to ; >---e--.!j x Figure 5.10 Position of mass cannot be observed and cannot be controlled using only velocity feedback.

232 CONTROL SYSTEM DESIGN or 9,=!i,/{3.'J,= (Il, --a)/ {3 which is the same as (6A.5). Note that the position and velocity gains {/, and {/" respectively, are proportional to the amounts we wish to move the coefficients from their open-loop positions. The position gain 9, is necessary to produce a stable system: 11, > O. But if the designer is willing to settle for a, = «, i.c. to accept the open-loop damping, then the gain {/, can be zero. This of course eliminatcs the need for a tachometer and reduces the hardware cost of the system. It is also possible to a lter the system damping without the use of a tachometer, by using an estimate w of the II ngula r velocity lij. This estimate is obtained by means of an observer as discussed in Chap. 7. Example 6B Stabilization of an inverted pendulum An inverted pendulum can readily be stabilized by a closed-loop feedback system, just, s a person of moderate dellterity can do il. A possible control system implementation is shown in Fig. 6.3, for (I pendulum constrained to rotate about a shaft at its bottom point. The acnnltor I a dc motor. The angular position of lht) pendulum, being equal to the position of the shaft to which it i II.lt3ched, is measured by mcan. of a poteotiometer. The!tllgular velocity in this case can ht measured by a "velocity pick-ou" nl the top of the pendulum. Sl.Ich It device could consist of n coil of wire Velocity pick-off u Figure 6.3 Implementation of system to stabilize inverted pendulum.

SHAPING THE DYNAMIC RESPONSE 233 in a magnetic field crc,)ted by a ~ m (l\l voltage in the ooil is proportional to the tinear velocity of the bob as it pa s~es permanent magnet in the pendulum bob. The induced the coil. And since the bop is at a fix ed distance from the pi v L point the linear ve locity is proportional La the angular veloci ty. TIle angular velocity could of course also be measured by means of a tachometer 011 the de molor shuft. As determined in Prob. 2.2, the dynamic equations governing the inverted pendulum in which the point of attachment does not translate is given by 8=w (6 3.1) w = 0 2 0 - (fw + {3u where a and {3 are given in Example 6A, with the inertia J being the total reflected inertia:.. where m is the pendulum bob mass and I is the distance of the bob from the pivot. The natural frequency 0 is given by 02=~ =--g-- J + me l + Jlml (Note that the motor inertia J,,, affects the natural frequency.) Since the linearization is valid only when the pendulum is nearly vertical, we shall assume that the control objective is to maintain IJ = O. Thus we have a si mple regulator problem. The matrices A and b for this problem are The open-loop characteristic polynomial is Is/ - AI = \ S - \ \ _0 2 S + a = S2 + as - 0 2 Thus a2 = _ 0 2 The open-loop system is unstable, of course. The controllability test matrix and the W matrix are given respectively by '(which are the same as they were for the instrument servo). And [(QW)T' = L~{3 1~{3] Thus the gain matrix required for pol e placement using (6.34), is Example 6C Control of spring-coupled masses The dynamics of a pair of spring-coupled masses, shown in Fig. 3.7(a), were shown in Example 3110 have the matrices lo l o 0 A = o 0 o 0 o o - KIM

234 CONTROL SYSTEM DESIGN Hence The system has the characteristic polynomial D(s)=s4+(KIM)s2 The controllability test and W matrices are given, respectively, by lo o Q = 0 I I 0 Multiplying we find that o -KIM QW = (QW)' = (QW)-I = l KIM o I] o 0 I 0 o I 0 0 I 0 0 0 (6C.1) (6C.2) (This rather simple result is not really as surprising as it may at first seem. Note that A is in the first companion form but using the right-to-left numbering convention. If the left-to-right numbering convention were used the A matrix would already be in the companion form of (6.11) and would not require transformation. The transformation matrix T given by (6C.2) has the effect of changing the state variable numbering order from left-to-right to right-to-ieft, and vice versa.) The gain matrix 9 is thus given by = [~ ~ ~ ~] [-2 - ii~ I Ml = [ : : _] goo 0 ii, if]. - K j M I 0 0 0 ii4 ii, A suitable pole "constellation" for the closed-loop process might be a Butterworth pattern as discussed in Sec. 6.5. To achieve this pattern the characteristic polynomial should be of the form Thus ii, = (1 + -/3)0 ii2 = (2 + -/3)0 2 ii, = (1 + /3)0' Thus the gain matrix 9 is given by 6.3 MULTIPLE-INPUT SYSTEMS If the dynamic system under consideration i = Ax + Bu