yield.pdf

Upload: amir-amiri

Post on 08-Aug-2018

212 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/22/2019 yield.pdf

    1/19

    CHAPTER

    7

    7.1 SECOND-ORDER SYSTEM

    Transfer Function

    This section introduces a basic system called a second-order system or a quadratic lag.

    Second-order systems are described by a second-order differential equation that relates

    the ouput variabley in Chap. 6. Some systems are inherently second-order, and they do not result from a

    series combination of two first-order systems. Inherently second-order systems are notextremely common in chemical engineering applications. Most second-order systems

    that we encounter will result from the addition of a controller to a first-order process.

    Lets examine an inherently second-order system and develop some terminology that

    will be useful in our analysis of the control of chemical processes.

    Consider a simple manometer as shown in Fig. 71. The pressure on both legs of

    the manometer is initially the same. The length of the fluid column in the manometer

    isL. At time t 0, a pressure difference is imposed across the legs of the manometer.

    Assuming the resulting flow in the manometer to be laminar and the steady-state fric-

    tion law for drag force in laminar flow to apply at each instant, we will determine the

    transfer function between the applied pressure difference P and the manometer read-

    ing h. If we perform a momentum balance on the fluid in the manometer, we arrive at

    the following terms:

    ( ) (Sum of for

  • 8/22/2019 yield.pdf

    2/19

    138 PART 2 LINEAR OPEN-LOOP SYSTEMS

    where

    Sum of forces

    causing fluid to move

    Un

    bbalanced pressure forces

    causing motion

    Frictional forces

    opposing motion

    Unbalanced pressure forcescausing motion

    ( ) P P D gh D1 22 2

    4 4p r p

    Frictional forces

    opposing motion

    Skin

    friction

    at wall

    Shear stress

    at wall

    Area in contact

    with wall

    Frictional forces

    opposing motionWal

    t ll pm

    pm

    pDLV

    DDL

    D

    dh

    dtDL( ) ( )

    8 8 1

    2(

    After (Final)Before (Initial)

    L

    h h/2

    h/2

    DReference level

    t= 0

    P1= 0 P1 P2P2= 0

    FIGURE 71

    Manometer.

  • 8/22/2019 yield.pdf

    3/19

    CHAPTER 7 HIGHER-ORDER SYSTEMS: SECOND-ORDER AND TRANSPORTATION LAG 139

    The rate of change of momentum of the fluid [the right side of Eq. (7.2)] may be

    expressed as

  • 8/22/2019 yield.pdf

    4/19

    140 PART 2 LINEAR OPEN-LOOP SYSTEMS

    2

    3

    162

    2 2

    1 2L

    g

    d h

    dt

    L

    D g

    dh

    dth

    P P

    g

    P

    g

    m

    r r r

    (7.4)

    (A more detailed version of the analysis of the manometer can be found in Bird et al.,

    1960). Note that as with first-order systems, standard form has a coefficient of 1 on the

    dependent variable term, h in this case. Second-order systems are described by a second-

    order differential equation. We may rewrite this Eq. (7.4) in general terms as

    t zt22

    22

    d Y

    dt

    dY

    dtY X t ( )

    (7.5)

    where

    t22

    3

    L

    g(7.6)

    216

    2zt

    m

    r

    L

    D g(7.7)

    X tP

    gY h( )

    r

    and

    (7.8)

    Solving for tand zfrom Eqs. (7.6) and (7.7) gives

    t 2

    3

    L

    gs

    (7.9)

    zm

    r

    8 3

    22D

    L

    gdimensionless

    (7.10)

    By definition, both tand zmust be positive. The reason for introducing tand zin theparticular form shown in Eq. (7.5) will become clear when we discuss the solution of

    Eq. (7.5) for particular forcing functionsX(t).

    Equation (7.5) is written in a standard form that is widely used in control theory.

    If the fluid column is motionless (dY/dt 0) and located at its rest position (Y 0)

    before the forcing function is applied, the Laplace transform of Eq. (7.4) becomes

    t zt2 2 2s Y s sY s Y s X s( ) ( ) ( ) ( )

    (7.11)

    From this, the transfer function follows:

    Y ( ) 1

  • 8/22/2019 yield.pdf

    5/19

    CHAPTER 7 HIGHER-ORDER SYSTEMS: SECOND-ORDER AND TRANSPORTATION LAG 141

    denominator oft2s2 2zts 1. All such systems are defined as second-order. Notethat it requires two parameters, tand z, to characterize the dynamics of a second-order

    system in contrast to only one parameter for a first-order system. We now discuss theresponse of a second-order system to some of the common forcing functions, namely,

    step, impulse, and sinusoidal.

    Step Response

    If the forcing function is a unit-step function, we have

    X s s( )

    1

    (7.13)

    In terms of the manometer shown in Fig. 71, this is equivalent to suddenly applying a

    pressure difference [such that X(t) P/rg 1] across the legs of the manometer attime t 0.

    Superposition will enable us to determine easily the response to a step function of

    any other magnitude.

    Combining Eq. (7.13) with the transfer function of Eq. (7.12) gives

    Y ss s s

    ( )

    1 12 12 2t zt

    (7.14)

    The quadratic term in this equation may be factored into two linear terms that contain

    the roots

    sa z

    t

    z

    t

    2 1

    (7.15)

    sb z

    t

    z

    t

    2 1

    (7.16)

    Equation (7.14) can now be written

    Y ss s s s sa b

    ( )

    1 2/t

    ( )( ) (7.17)

    The response of the system Y(t) can be found by inverting Eq. (7.17). The roots saand

    sb will be real or complex depending on value of the parameter z. The nature of the

    roots will, in turn, affect the form ofY(t). The problem may be divided into the threecases shown in Table 7.1. Each case will now be discussed.

    TABLE 71

    Step response of a second-order system

  • 8/22/2019 yield.pdf

    6/19

    142 PART 2 LINEAR OPEN-LOOP SYSTEMS

    CASE I STEP RESPONSE FOR y< 1. For this case, the inversion of Eq. (7.17) yields

    the result

    Y t ett( ) tan

    11

    1

    1

    2

    2 12

    zz

    t

    z

    zz t/ sin 1

    (7.18)

    To derive Eq. (7.18), use is made of the techniques of Chap. 3. Since z< 1, Eqs. (7.15)to (7.17) indicate a pair of complex conjugate roots in the left half-plane and a root at

    the origin. In terms of the symbols of Fig. 31, the complex roots correspond to s2 and

    s2* and the root at the origin to s6.The reader should realize that in Eq. (7.18), the argument of the sine function is in

    radians, as is the value of the inverse tangent term.

    By referring to Table 3.1, we see that Y(t) has the form

    Y t C e C t

    Ctt( ) 1 2

    23

    21 1z t zt

    zt

    / cos sin

    (7.19)

    The constants C1, C2, and C3 are found by partial fractions. The resulting equation is

    then put in the form of Eq. (7.18) by applying the trigonometric identity used in Chap. 4,

    Eq. (4.26). The details are left as an exercise for the reader. It is evident from Eq. (7.18)

    that Y(t) 1 as t.

    The nature of the response can be understood most clearly by plotting the solution

    to Eq. (7.17) as shown in Fig. 73, where Y(t) is plotted against the dimensionless vari-

    able t/t for several values ofz, including those above unity, which will be consideredin the next section. Note that for z< 1 all the response curves are oscillatory in nature

    and become less oscillatory as z is increased. The slope at the origin in Fig. 73 iszero for all values ofz. The response of a second-order system for z< 1 is said to beunderdamped.

    What is the physical significance of an underdamped response? Using the manom-

    eter as an example, if we step-change the pressure difference across an underdamped

    manometer, the liquid levels in the two legs will oscillate before stabilizing. The oscil-

    lations are characteristic of an underdamped response.

    CASE II STEP RESPONSE FOR y 1. For this case, the response is given by theexpression

    Y tt

    e t( ) / 1 1t

    t

    (7.20)

    This is derived as follows: Equations (7 15) and (7 16) show that the roots s and s are

  • 8/22/2019 yield.pdf

    7/19

  • 8/22/2019 yield.pdf

    8/19

    144 PART 2 LINEAR OPEN-LOOP SYSTEMS

    Using MATLAB/Simulink to Determine the Step Response of the Manometer

    Consider a manometer as illustrated in Fig. 71. The manometer is being used to determine the

    pressure difference between two instrument taps on an air line. The working fluid in the manometer

    is water. Determine the response of the manometer to a step change in pressure across the legs of

    the manometer.

    Data

    L 200 cm

    g 980 cm/s2

    Pg

    ttr

    0 010 0

    forcm for

    D 0.11 cm, 0.21 cm, 0.31 cm (Three Cases)

    Solution. From Eq. (7.4), we have the governing differential equation for the manometer:

    m

    r

    1 0 01

    1 0 3

    cP g/ cm s

    g/cmfor th

    . ( )

    .

    ee working fluid water,

    m

    r

    1 0 01

    1 0 3

    cP g/ cm s

    g/cmfor th

    . ( )

    .

    ee working fluid water,

    Actually, the response for z> 1 is not new. We saw it previously in the discussionof the step response of a system containing two first-order systems in series, for which

    the transfer function is

    Y s

    X s s s

    ( )

    ( )

    1

    1 11 2t t( )( ) (7.22)

    This is true for z> 1 because the roots s1 and s2 are real, and the denominator of Eq.(7.12) may be factored into two real linear factors. Therefore, Eq. (7.12) is equivalent to

    Eq. (7.22) in this case. By comparing the linear factors of the denominator of Eq. (7.12)

    with those of Eq. (7.22), it follows that

    t z z t 12 1 ( ) (7.23)

    t z z t 22 1 ( ) (7.24)

    Note that ift1t2, then tt1t2 and z 1. The reader should verify these results.

  • 8/22/2019 yield.pdf

    9/19

  • 8/22/2019 yield.pdf

    10/19

    146 PART 2 LINEAR OPEN-LOOP SYSTEMS

    The Simulink model for simulating the transfer functions is shown in Fig. 74, and the response is

    shown in Fig. 75.

    FIGURE 74

    Simulink diagram for manometer simulation.

    0.136s2 + 2.70s + 1

    0.136s2 + 0.738s + 1

    1

    Critically damped manometer

    Overdamped manometer

    Pressure forcingfunction

    10/s

    Scope

    1

    0.136s2 + 0.340s + 1

    1

    Underdamped manometer

    0

    2

    4

    6Height

    Underdamped

    Overdamped

    Criticallydamped

    8

    10

    12

  • 8/22/2019 yield.pdf

    11/19

  • 8/22/2019 yield.pdf

    12/19

  • 8/22/2019 yield.pdf

    13/19

  • 8/22/2019 yield.pdf

    14/19

  • 8/22/2019 yield.pdf

    15/19

    CHAPTER 7 HIGHER-ORDER SYSTEMS: SECOND-ORDER AND TRANSPORTATION LAG 151

    The presence ofs on the right side of Eq. (7.33) implies differentiation with respect to t

    in the time response. In other words, the inverse transform of Eq. (7.31) is

    Y td

    dtY t( ) ( ) impulse step ( ) (7.34)

    Application of Eq. (7.34) to Eq. (7.18) yields Eq. (7.32). This principle also yields the

    results for the next two cases.

    CASE II IMPULSE RESPONSE FOR y 1. For the critically damped case, the response

    is given by

    Y t te t( ) / 12t

    t

    (7.35)

    which is plotted in Fig. 78.

    CASE III IMPULSE RESPONSE FOR y> 1. For the overdamped case, the response is

    given by

    Y t ett( ) /

    1 1

    11

    2

    2

    t zz

    tz t sinh

    (7.36)

    which is also plotted in Fig. 78.

    To summarize, the impulseresponse curves of Fig. 78 show the same general

    behavior as the step response curves of Fig. 73. However, the impulse response always

    returns to zero. Terms such as decay ratio, period of oscillation, etc., may also be used

    to describe the impulse response. Many control systems exhibit transient responsessuch as those of Fig. 78.

    Sinusoidal Response

    If the forcing function applied to the second-order system is sinusoidal

    X t A t( ) sin w

    then it follows from Eqs. (7.12) and (4.23) that

    Y sA

    s s s( )

    w

    w t zt 2 2 2 2 2 1( )( ) (7.37)

    The inversion of Eq. (7.37) may be accomplished by first factoring the two quadratic

  • 8/22/2019 yield.pdf

    16/19

  • 8/22/2019 yield.pdf

    17/19

    154 PART 2 LINEAR OPEN-LOOP SYSTEMS

    units of time. The transportation lag parameter tis simply the time needed for a particleof fluid to flow from the entrance of the tube to the exit, and it can be calculated from

    the expression

    t volume of tube

    volumetric flow rate

    or t AL

    q

    (7.43)

    It can be seen from Fig. 710 that the relationship betweeny(t) andx(t) is

    y t x t( ) t( )

    (7.44)

    Subtracting Eq. (7.42) from Eq. (7.44) and introducing the deviation variablesXxxsand Yyysgive

    Y t X t ( ) t( )

    (7.45)

    If the Laplace transform ofX(t) isX(s), then the Laplace transform ofX(tt) is e

    stX(s).

    This result follows from the theorem on translation of a function, which was discussed

    in App. 3A. Equation (7.45) becomes

    Y s e X ss( ) ( ) t

    orY s

    X se s

    ( )

    ( )

    t

    (7.46)

    Therefore, the transfer function of a transportation lag is est.

    The transportation lag is quite common in the chemical process industries where

    a fluid is transported through a pipe. We shall see in a later chapter that the presence of

    a transportation lag in a control system can make it much more difficult to control. In

    general, such lags should be avoided if possible by placing equipment close together.

    They can seldom be entirely eliminated.

    APPROXIMATION OF TRANSPORT LAG. The transport lag is quite different from the

    other transfer functions (first-order, second-order, etc.) that we have discussed in that it

    is not a rational function (i.e., a ratio of polynomials.) As shown in Chap. 13, a system

    containing a transport lag cannot be analyzed for stability by the Routh test. The trans-

    port lag can also be difficult to simulate by computer. For these reasons, several approx-

    imations of transport lag that are useful in control calculations are presented here

  • 8/22/2019 yield.pdf

    18/19

    CHAPTER 7 HIGHER-ORDER SYSTEMS: SECOND-ORDER AND TRANSPORTATION LAG 155

    Keeping only the first two terms in the denominator gives

    e s

    s

    t

    t1

    1 (7.47)

    This approximation, which is simply a first-order lag, is a crude approximation of a

    transport lag. An improvement can be made by expressing the transport lag as

    ee

    e

    ss

    s

    t

    t

    t

    /

    /

    2

    2

    Expanding numerator and denominator in a Taylor series and keeping only terms of

    first-order give

    es

    s

    s

    t t

    t

    1 2

    1 2

    /

    /first-order Pad

    (7.48)

    This expression is also known as afirst-order Padapproximation.

    Another well-known approximation for a transport lag is the second-order Pad

    approximation:

    es s

    s s

    s

    t t t

    t t

    1 2 12

    1 2 12

    2 2

    2 2

    / /

    / / second-ordeer Pad(7.49)

    Equation (7.48) is not merely the ratio of two Taylor series; it has been optimized to

    give a better approximation.

    The step responses

    of the three approximations

    of transport lag presented

    here are shown in Fig. 711.

    The step response of e

    ts isalso shown for comparison.

    Notice that the response for

    the first-order Pad approxi-

    mation drops to1 before ris-

    ing exponentially toward 1.

    The response for the second-

    order Pad approximation

    jumps to

    1 and then descendsto below 0 before returning

    gradually back to1.

    Although none of the

    approximations for ets is

    very accurate, the approxima-

    i f t s i f l h i i l i li d b l fi d d d

    FIGURE 711

    Step response to approximation of the transport lag ets:

    (1)1

    1ts ; (2) first-order Pad; (3) second-order Pad ; (4) ets.

    01.0

    0.6

    0.2

    0.2

    0Y

    0.6

    1.0

    (2)

    (3)

    (1)

    (4)

    1 2t/

  • 8/22/2019 yield.pdf

    19/19