12
IN 227 Control Systems Design Lecture 10 Instructor: G R Jayanth Department of Instrumentation and Applied Physics Ph: 22933197 E-mail: [email protected]

Lecture 10 Notes

  • Upload
    abhinav

  • View
    215

  • Download
    1

Embed Size (px)

DESCRIPTION

control system design

Citation preview

  • IN 227 Control Systems Design

    Lecture 10

    Instructor: G R Jayanth

    Department of Instrumentation and Applied Physics

    Ph: 22933197

    E-mail: [email protected]

  • Motivation for plotting the root locus

    In the last couple of lectures, we learnt how the frequency domain representation of the open-loop transfer functions of the controller and the plant can be used for purposes of design of closed-loop control systems.

    Frequency domain design using Bode plots is both powerful and intuitive-it allows us to achieve the desired dynamic response for the overall system (say to a step input) in addition to the desired input tracking/ disturbance rejection properties.

    The dynamic response of a plant is decided primarily by its dominant poles, whose locations can be inferred approximately by using Bode plots. However, one of the (few) drawbacks of frequency domain based design is that it does not reveal the location of all the closed-loop poles. It would be advantageous to have a tool that enables direct visualization of the trajectory of all the closed-loop poles when our controller/ plant characteristics change so that we can ensure that the closed-loop performance variation is tolerable even when the open-loop characteristics change.

    Root locus plots do precisely that: they are plots of the trajectory of the closed-loop poles of a system with a given open-loop transfer function, when one of the parameters of the open-loop system varies. However, their application extends beyond just studying the changes in closed-loop pole locations when the open-loop system changes: they can also be used as design tools.

    In one sense, root locus plots are an improvement over Nyquist stability theory (and chronologically too, was proposed by W. R. Evans a couple of decades after H. Nyquist proposed his theory)-Nyquist stability theory seeks to verify if a particular region in the s-plane has a closed-loop pole or not. However, Root locus plots directly specify all the different regions in the s-plane where the closed-loop poles can possibly lie.

  • The root locus Consider an open-loop system possessing a single variable parameter K and open-loop transfer function G(s).

    The closed loop system is given by- Gcl(s)=KG(s)/(1+KG(s)). The root locus seeks to find all the regions in the s-plane where the poles of Gcl(s) can lie, i.e., all possible locations of the roots of 1+KG(s)=0 when the gain K changes from (typically) 0 to infinity.

    To find this, we note that if s0 is a root of 1+KG(s)=0, then G(s0)=-1/K. G(s0) is a complex number. Since K is variable, nothing definite can be said about the magnitude of G(s0). However, we can say something definite about its phase. The right-side of the equation (-1/K) can be expressed as (1/K)e-i. Thus, we see that the phase of G(s0) is fixed, and equal to either or its cyclic equivalents: -2l (l=0,1,2).

    Thus for a particular complex number s to be a root of 1+KG(s)=0, the necessary and sufficient condition is that

    . As a consequence, an equivalent but mathematically more useful definition emerges for the root locus: the root locus is the set of points in the s-plane for which .

    Significance of the parameter K: From the way that the closed-loop transfer function has been written, viz., Gcl(s)=KG(s)/(1+KG(s)), we get the impression that K should represent some gain of the controller that we can tune. However, root locus techniques can also be used to study the effect of variation in any other parameter (such as the location of poles and zeros of G(s)) by suitable rearrangement of the denominator polynomial.

    Example: Let us choose . Suppose there did indeed exist a gain K which we could tune, then the root locus studies the variation of the zeros of . However, if the gain K is a constant (say equal to 1), but the pole at -1 varies, the effect of the variation of this pole can be studied by looking at the zeros of (where p is nominally =1). If a particular s is a root of the numerator polynomial, this equation is not in a form where we can make definite predictions about the phase of G(s). However, when we rearrange, we see that the closed-loop transfer function can be written as

    Now we see that it is possible to study the variation of the pole p by plotting the locus of points where the phase of equals . Notice that G1(s) looks very different from G(s) and has nothing to do with the dynamics of the actual open-loop system. It is important to note that in each case, the parameter whose effect has to be studied has been brought out and the denominator has been expressed in the form 1+KG(s)=0, where G(s) does not contain the parameter whose variation we wish to study.

    Needless to say, root-loci study the effect of changes in just one parameter at a time.

    ( ) 2G s

    ( ) 2G s

    ( ) 1 [( 1)( 2)]G s s s

    1 [( 1)( 2)] 0K s s

    1 1 [( )( 2)] 0s p s

    2 2

    1 1( ) [ ]

    ( 2 1) 1 ( 2) ( 2 1)clG s

    s s p s s s

    2

    1( ) ( 2) ( 2 1)G s s s s 2

  • k

    -pk -zl

    Guidelines for drawing root-loci- A review

    Geometric interpretation of G(s): Suppose . We see that

    for any general point s in the s-plane, s+pk (k=1..n) and s+zl (l=1..m) are complex numbers shown on the right. Thus G(s) is the product (or ratio) of all such complex numbers. Thus, the magnitude of G(s) is the product of the magnitudes of each of the complex numbers in the numerator divided by the product of the magnitudes of complex numbers in the denominator. Likewise, the phase of G(s) is the sum of the phases of each of the numerator complex numbers and the difference of the phases of the denominator complex numbers: . This is useful in drawing the root loci.

    It is assumed in these lectures that students are already familiar with drawing root-loci. Thus, only the important properties are reviewed here. For more detailed guidelines, any standard textbook in Control Systems ,such as the reference indicated at the end, can be used.

    Rule 0: In general, if the closed-loop denominator polynomial is of nth degree, it has n distinct roots. As the parameter K varies, each of these roots trace out their own trajectory in the s-plane. Thus. there are as many branches to Root-locus as the degree of the denominator polynomial. Furthermore, since complex roots appear in conjugate pairs, the root locus is symmetric about the real-axis.

    Rule 1: Gcl(s)=KG(s)/(1+KG(s))=KN(s)/(D(s)+KN(s)), where G(s)=N(s)/D(s). Thus, as K0, the denominator polynomial tends to D(s). Thus, the branches of root locus start at the poles of the open-loop system. Likewise, as K, the denominator would be dominated by N(s). Thus, some of the branches of the root locus end in the zeros of the open-loop system.

    Rule 2: One the real axis, the regions of existence of the root locus can deduced from the geometric interpretation of G(s). They exist in regions which have an odd number of poles+zeros to their right.

    1

    1

    ( )..( )..( )( )

    ( )..( )..( )

    l m

    k n

    s z s z s zG s

    s p s p s p

    s

    ks pls z

    l

    1 1

    ( )m n

    l kG s

  • -pk -zl

    Rule 3: Those branches of root-locus that do not end in an open-loop zero go towards infinity as K. Thus, for an open-loop transfer-function whose numerator is of degree m and denominator is of degree n, there are n-m branches that tend to infinity. We now need to find where they exist: If we sit on one of those branches very far away from the origin, all the open-loop zeros and poles appear bunched together, i.e., they all appear to subtend the same angle at the far away location (schematic on the right). The algebraic sum of the angles subtended by poles and zeros is given by (n-m). For the point s to lie on the root locus, we require (n-m)= -2l , i.e., = (2l) /(n-m). Thus we see that, very far away from the origin, the branches are straight lines inclined at angles . It can be shown that all these straight lines intersect at a point on the real-axis given by

    Rule 4: The points on the real-axis at which the root loci break away into the complex plane of break in from the complex plane are obtained by solving the algebraic equation: where, G(s)=N(s)/D(s).

    Guidelines for drawing root-loci- A review

    s

    ks p

    ls z

    Far away point s

    0 ( ) ( )i in m

    s p z n m

    ( ) ( )( ) ( ) 0

    dN s dD sD s N s

    ds ds

  • Root locus as a tool for design

    If the root-locus plots were used just to study the effect of variation of a particular parameter on a closed-loop system, it would just be an analysis tool. However, it can also be used for design. The goal of this lecture is to discuss the use of root-locus in designing a control system.

    Example 1: Consider a plant given by P(s)=1/[(s+1)(s+2)(s+3)]. The plant is controlled by a proportional controller K. We require the dominant closed-loop poles to have either the desired damping or the desired natural frequency n.

    We first draw the root-locus of the plant. This is shown below. Since the plant is of 3rd order with no zeros, we see that the root locus has three branches going towards infinity at 60 and -180 respectively. The locus of poles in the s-plane that have constant damping is a straight line tilted at an angle -cos-1() to the real axis. We find the point P1 at which this line intersects the root-locus. This point and its complex conjugate represent the poles of the closed-loop system. The gain K is calculated by evaluating P(s) at the point P1.

    To find the gain at which the closed-loop system will have the desired natural frequency, we see that the locus of poles in the s-plane that have constant n is a circle of radius n. Therefore, we find the point P2 at which this circle intersects the root-locus. This point and its complex conjugate represent the poles of the closed-loop system. The gain K is calculated by evaluating P(s) at the point P2.

    Notice that since our system is of third order, we would have another closed-loop pole for each case above. They would be situated on the third branch of the locus, at points P10 and P20 respectively as shown below.

    -8 -7 -6 -5 -4 -3 -2 -1 0 1-5

    -4

    -3

    -2

    -1

    0

    1

    2

    3

    4

    5Root Locus

    Real Axis

    Imagin

    ary

    Axis

    =constant

    n=constant

    Root locus of P(s)

    P1

    P2

    P10 P20

  • Example 2: In example 1, we see that we cannot simultaneously choose the desired damping ratio and the desired natural frequency. Suppose, however, we want our closed-loop dominant poles to be at a specified location, then our controller should be designed to ensure that the root-locus passes through this location. In this example, we assume that the desired dominant closed-loop poles should be at the points -1.52i.

    Since a root-locus is symmetric about the real-axis, we consider only one of the desired roots, namely, -1.5+2i. We note that the root-locus of the plant does not pass through -1.5+2i. This is because the phase of P(s) at s= -1.5+2i is not -180. Instead it is -233. However, the root-locus will pass through -1.5+2i if the controller can somehow force the phase of the open-loop transfer function C(j)P(j) to be -180 at -1.5+2i. Thus, our controller should be designed to add a phase of 53.

    Any transfer-function whose phase is 53 at -1.5+2i can potentially serve as our controller. However, we seek simplicity in the design of the controller, because simplicity promotes understanding.

    The simplest controller that can add a phase of 53 is of the form C(s)=s+z0, where the zero z0 subtends an angle 53 at

    -1.5+2i. This would imply that z0=3. Since this controller is not causal it cannot unfortunately be realized. However it does give us a clue: if our controller is of the form K(s+z0)/(s+p0), where the pole p0 is placed very far back along the imaginary axis, then the controller can not only be realized but will also adds approximately 53 at -1.5+2i , since the dominant contribution is only from z0. Thus, we choose (arbitrarily) that p0=11. For this choice, p0 adds a phase -12 at -1.5+2i. Thus, the zero z0 should add a phase of 53+12=65 for the root-locus to pass through -1.5+2i. Using simple trigonometry, this places the zero at -2.44. Finally, the gain K is calculated by evaluating |C(s)P(s)| at s= -1.5+2i. |C(s)P(s)|=0.021. Thus, K=1/ |C(s)P(s)| =46.67. Thus our final controller is C(s)=46.67(s+2.44)/(s+11). This is a lead compensator. The root locus, plotted below, is now seen to pass through the desired points.

    Root locus as a tool for design

    ~ 53z

    -z0 -p0

    ~ 0p

    1.5 2i

    -10 -8 -6 -4 -2 0

    -6

    -4

    -2

    0

    2

    4

    6

    Root Locus

    Real Axis

    Imagin

    ary

    Axis 1.5 2i

    Desired characteristics of the

    controller transfer function

  • -6 -4 -2 0

    -10

    -8

    -6

    -4

    -2

    0

    2

    4

    6

    8

    10

    Root Locus

    Real Axis

    Imagin

    ary

    Axis

    Example 3: Example 2 clearly illustrated the use of root-locus in designing the location of dominant poles: We compute the phase of G(s) at the desired location of the closed-loop pole and find a controller C(s) that supplies the deficit phase to ensure that the total phase of the open-loop system is -180 (360l). The gain K of the controller is finally calculated using the magnitude criterion, namely |KC(s)P(s)| =1. In example 2 just a single dominant zero was used to supply the requisite phase. However, if the deficit is very large we may need to use two or more zeros (and correspondingly add two far away poles in order to realize a causal system). Thus, for example, if we desired our closed-loop dominant pole to be at -33i, we note that the total phase of G(s) at -3+3i is -322, which is -142 away from -180.

    If a single zero has to supply this phase, it has to be situated well into the right-half plane. This is undesirable because this would guarantee that the root-locus would exist along the real-axis in the right-half plane and thus ensure that one of the poles may be unstable. Thus, we choose this to be supplied by two zeros, both in the left half plane. We choose one to supply +90 and the other to supply the reminder, viz., +52 .

    This places the first zero at -3 and the second zero at -5.34. To realize a causal controller, we need add two poles. We choose to add these poles far away, at -30. The total phase added by these two poles at -3+3i is about 13. To compensate, we correct the position of the second zero so that it subtends 52+13=65 at -3+3i . Thus, in the second iteration, we obtain its position to be at -4.4. Thus, the controller is of the form C(s)=K(s+3)(s+4.4)/(s+30)2. The gain K is found by setting |KC(s)P(s)| =1. This gives us K=2542, so that the controller is given by C(s)=2542(s+3)(s+4.4)/(s+30)2. The root locus is shown below.

    Root locus as a tool for design

    3 3i

  • Example 4: We see that root-locus based design is excellent for designing compensators that decide the dominant dynamics. However, as control engineers, our prime area of interest is not the dominant dynamics but the low-frequency region (low relative to the poles of the plant), where we want to track a specified input or reject some disturbance. The low frequency gain of an open-loop system is given by . In the schematic on the right we see that . Generally we would like the low frequency gain to be adequately high. If the gain is not high enough, we can add a pole p0 close to the origin (i.e., p0
  • When we examine what we focused the most on while designing in frequency domain and the root locus, we notice that we spent the maximum time achieving high gain in frequency domain-based design. Any attention we paid to phase was done at the end of the design process where we used zero-placement to improve the phase-margin. In contrast, root-locus based design focused a lot on placement of the dominant poles and relatively little on achieving high gain.

    However, since control engineers generally ought to focus on high gain much more than dominant dynamics, frequency domain design tools are preferable over root-locus. Nevertheless, the two complement one another: we can first design the system for low frequency control performance using Bode plots and then shift to root-locus to achieve the desired dominant dynamics by placing the closed-loop poles at desired locations.

    Unlike the frequency domain magnitude and phase plots of the controller and the plant that add up, root-loci of the controller and the plant do not add up. Indeed, the root-locus of the combined open-loop system could be dramatically different from that of each individual component. Thus, root-locus demands that we invest much more effort in building our intuition about the characteristics of root-loci before we use them as design tools.

    When we have such an excellent and intuitive tool as Bode plot for the design of one degree of freedom (1-DOF) control systems it is hard to think of why one may want to invest extra energy and learn to design a control system using root-locus. However, we shall see in Lecture 11 that this is not the case for 2-DOF control systems root locus offers key insights for design of robust control systems that are not as intuitively apparent in the case Bode plots for 2DOF control systems.

    Control design using root locus vis--vis Bode plots: some comments

  • While our opinion on root-locus as a design tool for 1-DOF control systems was not wholly enthusiastic, the same cannot be said about its virtues as a tool for analysis. For example, in the previous lecture, we designed a PI-based control system whose step response has been shown on the right (the blue trace). We notice a long tail in the response, i.e., after quickly reaching a value close to the reference, the closed-loop systems response is dominated by that of a very slow pole. There is simply no way for us to directly anticipate that closed-loop response would have such a slow pole simply by looking at the Bode plot of the open-loop system.

    However, if we turn to root locus, we can immediately see why such a response can occur: we notice that if we employ PI-control, a segment of the root-locus exists between the pole and the zero of the PI-controller. Thus, one of the roots of the closed-loop system has to lie on this segment. Since z0, p0 ~0 , this is a very slow pole. Now, if the open-loop gain is sufficiently high, the closed-loop pole will lie very close to the open-loop zero and the effect of the two on the closed-loop response will nearly cancel each other. However, if the gain is insufficient, the closed-loop pole is somewhere midway between the two ends of the segment and can contribute substantially to the overall response. It is this situation that we find in the PI-control that was designed in the last lecture.

    Thus, even if we prefer to design a controller using frequency-domain techniques, it is always prudent to look at the root-locus of the open-loop system later and identify the possibilities for maladies such as the one above.

    Root locus as a tool for analysis

    0 0.5 1 1.5 2 2.50

    0.2

    0.4

    0.6

    0.8

    1

    1.2

    1.4Step Response

    Time (sec)

    Am

    plit

    ude

    Step response of PI-

    controlled closed-loop

    system

    Long tail

    -1 -0.5 0

    -0.8

    -0.6

    -0.4

    -0.2

    0

    0.2

    0.4

    0.6

    0.8

    1

    1.2

    Root Locus

    Real Axis

    Imagin

    ary

    Axis

    0z 0p

    Segment of root-locus

  • References

    (1) G. F. Franklin, J. David Powell and A. Emami-Naeini, Feedback Control of Dynamic Systems, Pearson Education (2002)

    (2) K Ogata, Modern Control Engineering, Pearson (2003)