Sie sind auf Seite 1von 19

Confirming Pages

137
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 variable y to the input variable x (the forcing function) with time as the inde-
pendent variable.

A
d y
dt
B
dy
dt
Cy x t
2
2
( )

(7.1)
A second-order system can arise from two first-order systems in series, as we saw
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 not
extremely 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
is L. 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:

( ) ( Sumof forces causing fluid to move Rate of change of momentumof fluid)

(7.2)

HIGHER-ORDER SYSTEMS:
SECOND-ORDER AND
TRANSPORTATION LAG
cou9789x_ch07_137-162.indd 137 8/14/08 5:58:48 PM
Confirming Pages
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 forces
causing motion

__
,

( ) P P
D
gh
D
1 2
2 2
4 4
p
r
p


Frictional forces
opposing motion
Skin

_
,


friction
at wall
Shear stress
at wall

_
,

_
,

_
,

Area in contact
with wall


Frictional forces
opposing motion
Wal

_
,

t
ll
p
m
p
m
p DL
V
D
DL
D
dh
dt
DL ( ) ( )

_
,

_
,


8 8 1
2
( ))

The term for the skin friction at the wall is obtained from the Hagen-Poiseuille relation-
ship for laminar flow (McCabe, Smith and Harriott, 2004). Note that V is the average
velocity of the fluid in the tube, which is also the velocity of the interface, which is
equal to
1
2
dh dt / (see Fig. 72 ).
After (Final) Before (Initial)
L
h h/2
h/2
D
Reference level
t 0
P
1
0 P
1
P
2
P
2
0
FIGURE 71
Manometer.
cou9789x_ch07_137-162.indd 138 8/14/08 5:58:50 PM
Confirming Pages
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

( ) ( Rate of change of momentum mass veloc
d
dt
iity momentumcorrection factor

)
r
p D
L
2
4

_
,,

( )

_
,

( )

_
,

b
r
p
b
dV
dt
D
L
d h
dt

2 2
2
4
1
2

The momentum correction factor b accounts for the fact that the fluid has a parabolic
velocity profile in the tube, and the momentum must be expressed as b m V for laminar
flow (see McCabe, Smith and Harriott, 2004). The value of b for laminar flow is 4/3. Sub-
stituting the appropriate terms into Eq. (7.2) produces the desired force balance equation
for the manometer.

r
p p D
L
d h
dt
P P
2 2
2
1 2
4
4
3
1
2

_
,

_
,

_
,

( )
DD
gh
D
D
dh
dt
DL
2 2
4 4
8 1
2
r
p m
p

_
,

_
,

(

(7.3)

Rearranging Eq. (7.3), we obtain

r
p m D
L
d h
dt D
2 2
2
4
4
3
1
2
8

_
,

_
,

_
,

_
,

_
,

( ) ( )
1
2 4 4
2
1 2
2
dh
dt
DL gh
D
P P
D
p r
p p


and finally, dividing both sides by r g ( p D
2
/4), we arrive at the standard form for a
second-order system.
FIGURE 72
Average velocity of the fluid in the manometer.
h
V
V
h/2
h/2
Reference level
t 0
P
1
P
2
cou9789x_ch07_137-162.indd 139 8/14/08 5:58:53 PM
Confirming Pages
140 PART 2 LINEAR OPEN-LOOP SYSTEMS

2
3
16
2
2 2
1 2
L
g
d h
dt
L
D g
dh
dt
h
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 zt
2
2
2
2
d Y
dt
dY
dt
Y X t ( )

(7.5)

where

t
2
2
3

L
g
(7.6)

2
16
2
zt
m
r

L
D g
(7.7)


X t
P
g
Y h ( )

r
and

(7.8)
Solving for t and z from Eqs. (7.6) and (7.7) gives

t
2
3
L
g
s

(7.9)

z
m
r

8 3
2
2
D
L
g
dimensionless

(7.10)
By definition, both t and z must be positive. The reason for introducing t and z in the
particular form shown in Eq. (7.5) will become clear when we discuss the solution of
Eq. (7.5) for particular forcing functions X ( 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 zt
2 2
2 s Y s sY s Y s X s ( ) ( ) ( ) ( )

(7.11)
From this, the transfer function follows:

Y s
X s s s
( )
( )


1
2 1
2 2
t zt

(7.12)
The transfer function given by Eq. (7.12) is written in standard form, and we will show
later that other physical systems can be represented by a transfer function having the
cou9789x_ch07_137-162.indd 140 8/14/08 5:58:54 PM
Confirming Pages
CHAPTER 7 HIGHER-ORDER SYSTEMS: SECOND-ORDER AND TRANSPORTATION LAG 141
denominator of t
2
s
2
2 z t s 1. All such systems are defined as second-order. Note
that it requires two parameters, t and z , to characterize the dynamics of a second-order
system in contrast to only one parameter for a first-order system. We now discuss the
response 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 / r g 1] across the legs of the manometer at
time 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 s
s s s
( )

1 1
2 1
2 2
t zt

(7.14)
The quadratic term in this equation may be factored into two linear terms that contain
the roots

s
a

z
t
z
t
2
1

(7.15)

s
b

z
t
z
t
2
1

(7.16)
Equation (7.14) can now be written

Y s
s s s s s
a b
( )

1
2
/t
( )( )
(7.17)
The response of the system Y ( t ) can be found by inverting Eq. (7.17). The roots s
a
and
s
b
will be real or complex depending on value of the parameter z . The nature of the
roots will, in turn, affect the form of Y ( t ). The problem may be divided into the three
cases shown in Table 7.1 . Each case will now be discussed.
TABLE 71
Step response of a second-order system
Case y Nature of roots Description of response
I < 1 Complex Underdamped or oscillatory
II 1 Real and equal Critically damped
III > 1 Real Overdamped or nonoscillatory
cou9789x_ch07_137-162.indd 141 8/14/08 5:58:55 PM
Confirming Pages
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 e
t
t
( ) tan




1
1
1
1
2
2 1
2
z
z
t
z
z
z 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 s
2
and
s
2
*
and the root at the origin to s
6
.
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
C
t
t
( )

1 2
2
3
2
1 1
z t
z
t
z
t
/
cos sin

_
,


(7.19)
The constants C
1
, C
2
, and C
3
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 of z , including those above unity, which will be considered
in 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 is
zero for all values of z . The response of a second-order system for z < 1 is said to be
underdamped.
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. la1id to b6(4.26 1 TTj0 Td( )Tj-0.004 Tc 0.0974 Tw T*(The constants )Tj/T1_3 1 Tf0 Tc 0 Tw 5.916 0 Td(C)Tj/T1_0 1 Tf-0.se. )S.oty, whY(BTw 5.77 404.68 Tm( )Tj-0.004 nse of a second- includidamped resp4 nse of a second- in1s0.25 Tc ( 9 )Tj/T1_)Tj/T1_3 1 Tf1.75 0 Td(Y)Tj/T1_0 1 Tf[( )25Tw 0e The resulting/Tecondl534 The cre0.00u3e-871und/T1_0hd
Confirming Pages
CHAPTER 7 HIGHER-ORDER SYSTEMS: SECOND-ORDER AND TRANSPORTATION LAG 145
In terms of transformed deviation variables, this becomes

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

wher e

Y h h X
P P
L
g D
s
s


and
g g
and

r r
t z
m
r

_
,

3
2
8
2
33
2
L
g

Lets calculate the time constant for the manometer.

t
2
3
2 200
3 980
0 369
2
L
g
( )
( )
.
cm
cm /s
s

and the damping coefficient for the three different tube diameters

z
m
r

8 3
2
8 0 01
1 0
2 3
D
L
g D
[ . ( )]
( . )(
g/ cm s
g/cm
22 2 2
3 200
2 980
0 0443
)
( )
( )
. cm
cm/s

D

Diameter (cm)
0.11 3.66
0.21 1.00
0.31 0.46

Clearly we have one underdamped system ( z < 1), one critically damped system ( z 1), and
one overdamped system ( z > 1). One method of obtaining the responses is to substitute the values
of t and z into Eqs. (7.18), (7.20), and (7.21) and plot the resulting equations, realizing that the forc-
ing function is 10 times a unit step. Another way to obtain the responses is to use MATLAB and
Simulink to obtain the response of the transfer function Y/X to the forcing function input X.

Y s
X s s s
X
s
( )
( )

1
2 1
10
2 2
t zt
and

The three necessary transfer functions are as follows:
Diameter (cm) s z s
2
2zs Transfer function
0.11 0.369 3.66 0.136 2.70
1
0 136 2 70 1
2
. . s s
0.21 0.369 1.00 0.136 0.738
1
0 136 0 738 1
2
. . s s
0.31 0.369 0.46 0.136 0.340
1
0 136 0 340 1
2
. . s s
(continued)
cou9789x_ch07_137-162.indd 145 8/14/08 5:58:58 PM
Confirming Pages
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.
0.136s
2
+ 2.70s + 1
0.136s
2
+ 0.738s + 1
1
Critically damped manometer
Overdamped manometer
Pressure forcing
function
10/s
Scope
1
0.136s
2
+ 0.340s + 1
1
Underdamped manometer
FIGURE 75
Manometer response to step input.
Time
0
0
2
4
6 Height
Underdamped
Overdamped
Critically
damped
8
10
12
1 2 3 4 5 6 7 8 9 10
cou9789x_ch07_137-162.indd 146 8/14/08 5:58:59 PM
Confirming Pages
CHAPTER 7 HIGHER-ORDER SYSTEMS: SECOND-ORDER AND TRANSPORTATION LAG 147
Substituting the values for t and z into Eqs. (7.18), (7.20), and (7.21), we get

Y t e
t
t
( ) 10 1
1
1




z
z
t
z
z
z t
2
2 1
2
1
1
/
sin tan

_
,

1
]
1
1
underdamped manometer
10 1 1.1 33
1.93
1.09 rad
e t
t

1 25
1
2 41
.
sin .
tan

_
,

1
]
1
1

_
,

1
]
1
Y t
t
e
t
( ) 10 1 1
/


t
t


10
0 369
0 369
1 1 critical
t
e
t
.
/ .

_
,

1
]
1
lly damped manometer
( ) Y t e
t

10 1 1
2 z t
z
/
cosh
tt t
t
z
z
z
t

2
2
1
1 sinh

_
,

1
]
1
1
overdampeed manometer
1 cosh(9.54 ) 1.04 si

10
9 92
e t
t .
nnh(9.54 ) t [ ]


Plotting these responses gives the same results as the Simulink model.
On a practical note, notice that t increases with the total length of the fluid column and that z
increases with the viscosity of the fluid. If the damping coefficient z is small (< < 1.0), the response
of the manometer to a change in pressure can be very oscillatory, and it becomes difficult to obtain
accurate readings of the pressure. To dampen the oscillations, it is common practice to place a
restriction on the bend of the tube. This increases the drag force of the fluid and is equivalent to
increasing m in the equation for z . Such a restriction (a partially open valve) is called a snubber.
Terms Used to Describe an Underdamped System
Of these three cases, the underdamped response occurs most frequently in control sys-
tems. Hence a number of terms are used to describe the underdamped response quan-
titatively. Equations for some of these terms are listed below for future reference. In
general, the terms depend on z and/or t . All these equations can be derived from the
time response as given by Eq. (7.18); however, the mathematical derivations are left to
the reader as exercises.
1. Overshoot. Overshoot is a measure of how much the response exceeds the ultimate
value following a step change and is expressed as the ratio A/B in Fig. 76 .
The overshoot for a unit step is related to z by the expression

Overshoot exp

pz
z 1
2

_
,


(7.25)

This relation is plotted in Fig. 77 . The overshoot increases for decreasing z .
cou9789x_ch07_137-162.indd 147 8/14/08 5:59:01 PM
Confirming Pages
CHAPTER 7 HIGHER-ORDER SYSTEMS: SECOND-ORDER AND TRANSPORTATION LAG 151
The presence of s 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 t
d
dt
Y 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
( )
/


1
2
t
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 e
t
t
( )
/

1 1
1
1
2
2
t
z
z
t
z 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 responses
such 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 s
A
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
terms to give

Y s
A
s j s j s s s s
a b
( )

w t
w w
/
2
( )( )( )( )

(7.38)
cou9789x_ch07_137-162.indd 151 8/14/08 5:59:05 PM
Confirming Pages
154 PART 2 LINEAR OPEN-LOOP SYSTEMS
units of time. The transportation lag parameter t is simply the time needed for a particle
of 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 between y ( t ) and x ( t ) is

y t x t ( ) t ( )

(7.44)
Subtracting Eq. (7.42) from Eq. (7.44) and introducing the deviation variables X x x
s
and Y y y
s
give

Y t X t ( ) t ( )

(7.45)
If the Laplace transform of X ( t ) is X ( s ), then the Laplace transform of X ( t t ) is e


s t
X ( 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 s
s
( ) ( )
t

or

Y s
X s
e
s
( )
( )

t

(7.46)

Therefore, the transfer function of a transportation lag is e
s t
.
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.
One approach to approximating the transport lag is to write e
t s
as 1/ e
t s
and to
express the denominator as a Taylor series; the result is

e
e s s s
s
s



t
t
t t t
1 1
1 2 3
2 2 3 3
/ / !

cou9789x_ch07_137-162.indd 154 8/14/08 5:59:08 PM
Confirming Pages
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
t

1
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

e
e
e
s
s
s

t
t
t
/
/
2
2

Expanding numerator and denominator in a Taylor series and keeping only terms of
first-order give

e
s
s
s

t
t
t

1 2
1 2
/
/
first-order Pad

(7.48)
This expression is also known as a first-order Pad approximation.
Another well-known approximation for a transport lag is the second-order Pad
approximation:

e
s 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
t s
is
also shown for comparison.
Notice that the response for
the first-order Pad approxi-
mation drops to 1 before ris-
ing exponentially toward 1.
The response for the second-
order Pad approximation
jumps to 1 and then descends
to below 0 before returning
gradually back to 1.
Although none of the
approximations for e
t s
is
very accurate, the approxima-
tion for e
t s
is more useful when it is multiplied by several first-order or second-order
transfer functions. In this case, the other transfer functions filter out the high-frequency
content of the signals passing through the transport lag, with the result that the transport
lag approximation, when combined with other transfer functions, provides a satisfactory
result in many cases. The accuracy of a transport lag can be evaluated most clearly in
terms of frequency response, a topic covered later in this book.
FIGURE 711
Step response to approximation of the transport lag e
ts
:
(1)
1
1 ts
; (2) first-order Pad; (3) second-order Pad ; (4) e
ts
.
0
1.0
0.6
0.2
0.2
0 Y
0.6
1.0
(2)
(3)
(1)
(4)
1 2 t/
cou9789x_ch07_137-162.indd 155 8/14/08 5:59:12 PM

Das könnte Ihnen auch gefallen