Vous êtes sur la page 1sur 49

Control System Design

Based on Frequency
Response Analysis Pratik N Sheth
BITS Pilani Department of Chemical Engineering
Pilani Campus
Contents

Closed Loop Behavior


Bode Stability Criterion

Dr Pratik N Sheth, Dept of Chemical Engg, BITS Pilani, Pilani Campus


Closed Loop Behavior

In general, a feedback control system should satisfy the following


design objectives:
1. Closed-loop stability
2. Good disturbance rejection (without excessive control action)
3. Fast set-point tracking (without excessive control action)
4. A satisfactory degree of robustness to process variations and
model uncertainty
5. Low sensitivity to measurement noise

Dr Pratik N Sheth, Dept of Chemical Engg, BITS Pilani, Pilani Campus


Figure 14.1 Block diagram with a disturbance D and
measurement noise N.
Dr Pratik N Sheth, Dept of Chemical Engg, BITS Pilani, Pilani Campus
The block diagram of a general feedback control system is
shown in Fig. 14.1.
It contains three external input signals: set point Ysp, disturbance
D, and additive measurement noise, N.

Gd Gc G K mGcGvG p
Y = D N + Ysp (14-1)
1 + Gc G 1 + GcG 1 + GcG

Gd Gm Gm Km
E = D N + Ysp (14-2)
1 + GcG 1 + GcG 1 + GcG

Gd GmGc Gv GmGcGv K mGcGv


U = D N + Ysp (14-3)
1 + GcG 1 + GcG 1 + GcG

where G GvGpGm.
5
Example 14.1
Consider the feedback system in Fig. 14.1 and the following
transfer functions:
0.5
G p = Gd = , Gv = Gm = 1
1 2s
Suppose that controller Gc is designed to cancel the unstable
pole in Gp:
3 (1 2 s )
Gc =
s +1
Evaluate closed-loop stability and characterize the output
response for a sustained disturbance.

6
Solution
The characteristic equation, 1 + GcG = 0, becomes:
3 (1 2s) 0.5
1 + = 0
s + 1 1 2s
or
s + 2.5 = 0
In view of the single root at s = -2.5, it appears that the closed-
loop system is stable. However, if we consider Eq. 14-1 for
N = Ysp = 0,

Gd 0.5 ( s + 1)
Y = D = D
1 + GcG (1 2 s )( s + 2.5)

7
This transfer function has an unstable pole at s = +0.5. Thus,
the output response to a disturbance is unstable.
Furthermore, other transfer functions in (14-1) to (14-3) also
have unstable poles.
This apparent contradiction occurs because the characteristic
equation does not include all of the information, namely, the
unstable pole-zero cancellation.

Example 14.2
Suppose that Gd = Gp, Gm = Km and that Gc is designed so that the
closed-loop system is stable and |GGc | >> 1 over the frequency
range of interest. Evaluate this control system design strategy for
set-point changes, disturbances, and measurement noise. Also
consider the behavior of the manipulated variable, U.

8
Solution
Because |GGc | >> 1,
1 Gc G
0 and 1
1 + GcG 1 + Gc G

The first expression and (14-1) suggest that the output response
to disturbances will be very good because Y/D 0. Next, we
consider set-point responses. From Eq. 14-1,

Y K mGc Gv G p
=
Ysp 1 + GcG

Because Gm = Km, G = GvGpKm and the above equation can be


written as,
Y GcG
=
Ysp 1 + Gc G
9
For |GGc | >> 1,
Y
1
Ysp

Thus, ideal (instantaneous) set-point tracking would occur.


Choosing Gc so that |GGc| >> 1 also has an undesirable
consequence. The output Y becomes sensitive to noise because
Y - N (see the noise term in Eq. 14-1). Thus, a design tradeoff
is required.

Dr Pratik N Sheth, Dept of Chemical Engg, BITS Pilani, Pilani Campus


Bode Stability Criterion
Before considering the basis for the Bode stability criterion, it is
useful to review the General Stability Criterion of Section 11.1:
A feedback control system is stable if and only if all roots of the
characteristic equation lie to the left of the imaginary axis in the
complex plane.
The Bode stability criterion has two important advantages in
comparison with the Routh stability criterion of Chapter 11:
1. It provides exact results for processes with time delays, while
the Routh stability criterion provides only approximate results
due to the polynomial approximation that must be substituted
for the time delay.
2. The Bode stability criterion provides a measure of the relative
stability rather than merely a yes or no answer to the question,
Is the closed-loop system stable?
Dr Pratik N Sheth, Dept of Chemical Engg, BITS Pilani, Pilani Campus
Root analysis

Dr Pratik N Sheth, Dept of Chemical Engg, BITS Pilani, Pilani Campus


Imp Definitions

1. A critical frequency c is defined to be a value of for


which OL ( ) = 180o. This frequency is also referred to as
a phase crossover frequency.
2. A gain crossover frequency g is defined to be a value of
for which AROL ( ) = 1 .

Dr Pratik N Sheth, Dept of Chemical Engg, BITS Pilani, Pilani Campus


Bode Stability Criterion. Consider an open-loop transfer function
GOL=GcGvGpGm that is strictly proper (more poles than zeros) and
has no poles located on or to the right of the imaginary axis, with
the possible exception of a single pole at the origin. Assume that
the open-loop frequency response has only a single critical
frequency c and a single gain crossover frequency g. Then the
Chapter 14

closed-loop system is stable if AROL( c ) < 1. Otherwise it is


unstable.

Some of the important properties of the Bode stability criterion


are:
1. It provides a necessary and sufficient condition for closed-
loop stability based on the properties of the open-loop transfer
function.
2. Unlike the Routh stability criterion of Chapter 11, the Bode
stability criterion is applicable to systems that contain time
delays. 14
3. The Bode stability criterion is very useful for a wide range of
process control problems. However, for any GOL(s) that does
not satisfy the required conditions, the Nyquist stability
criterion of Section 14.3 can be applied.
4. For systems with multiple c or g , the Bode stability
Chapter 14

criterion has been modified by Hahn et al. (2001) to provide a


sufficient condition for stability.

In order to gain physical insight into why a sustained oscillation


occurs at the stability limit, consider the analogy of an adult
pushing a child on a swing.
The child swings in the same arc as long as the adult pushes at
the right time, and with the right amount of force.
Thus the desired sustained oscillation places requirements on
both timing (that is, phase) and applied force (that is,
amplitude).
15
By contrast, if either the force or the timing is not correct, the
desired swinging motion ceases, as the child will quickly
exclaim.
A similar requirement occurs when a person bounces a ball.
To further illustrate why feedback control can produce
Chapter 14

sustained oscillations, consider the following thought


experiment for the feedback control system in Figure 14.4.
Assume that the open-loop system is stable and that no
disturbances occur (D = 0).
Suppose that the set point is varied sinusoidally at the critical
frequency, ysp(t) = A sin(ct), for a long period of time.
Assume that during this period the measured output, ym, is
disconnected so that the feedback loop is broken before the
comparator.

16
Chapter 14

Figure 14.4 Sustained oscillation in a feedback control system.

17
After the initial transient dies out, ym will oscillate at the
excitation frequency c because the response of a linear system
to a sinusoidal input is a sinusoidal output at the same frequency
(see Section 13.2).
Suppose that two events occur simultaneously: (i) the set point
is set to zero and, (ii) ym is reconnected. If the feedback control
Chapter 14

system is marginally stable, the controlled variable y will then


exhibit a sustained sinusoidal oscillation with amplitude A and
frequency c.
To analyze why this special type of oscillation occurs only when
= c, note that the sinusoidal signal E in Fig. 14.4 passes
through transfer functions Gc, Gv, Gp, and Gm before returning to
the comparator.
In order to have a sustained oscillation after the feedback loop is
reconnected, signal Ym must have the same amplitude as E and a
-180 phase shift relative to E.
18
Note that the comparator also provides a -180 phase shift due
to its negative sign.
Consequently, after Ym passes through the comparator, it is in
phase with E and has the same amplitude, A.
Thus, the closed-loop system oscillates indefinitely after the
Chapter 14

feedback loop is closed because the conditions in Eqs. 14-7


and 14-8 are satisfied.
But what happens if Kc is increased by a small amount?
Then, AROL(c) is greater than one and the closed-loop system
becomes unstable.
In contrast, if Kc is reduced by a small amount, the oscillation
is damped and eventually dies out.

19
Example 14.3
A process has the third-order transfer function (time constant in
minutes),
2
G p( s) =
(0.5s + 1)3
Chapter 14

Also, Gv = 0.1 and Gm = 10. For a proportional controller, evaluate


the stability of the closed-loop control system using the Bode
stability criterion and three values of Kc: 1, 4, and 20.

Solution
For this example,

2 2K c
G OL = G cG vG pG m = ( K c )(0.1) 3
(10) =
(0.5s + 1) (0.5s + 1)3

20
Figure 14.5 shows a Bode plot of GOL for three values of Kc.
Note that all three cases have the same phase angle plot because
the phase lag of a proportional controller is zero for Kc > 0.
Next, we consider the amplitude ratio AROL for each value of Kc.
Based on Fig. 14.5, we make the following classifications:
Chapter 14

Kc AROL ( for = c ) Classification


1 0.25 Stable
4 1 Marginally stable
20 5 Unstable

21
Chapter 14

Figure 14.5 Bode plots for GOL = 2Kc/(0.5s+1)3.

22
Feedforward and Ratio Control
In Chapter 8 is was emphasized that feedback control is an
important technique that is widely used in the process industries.
Its main advantages are as follows.
Chapter 15

1. Corrective action occurs as soon as the controlled variable


deviates from the set point, regardless of the source and type
of disturbance.
2. Feedback control requires minimal knowledge about the
process to be controlled; it particular, a mathematical model
of the process is not required, although it can be very useful
for control system design.
3. The ubiquitous PID controller is both versatile and robust. If
process conditions change, retuning the controller usually
produces satisfactory control.
23
However, feedback control also has certain inherent
disadvantages:
1. No corrective action is taken until after a deviation in the
controlled variable occurs. Thus, perfect control, where the
controlled variable does not deviate from the set point during
Chapter 15

disturbance or set-point changes, is theoretically impossible.


2. Feedback control does not provide predictive control action
to compensate for the effects of known or measurable
disturbances.
3. It may not be satisfactory for processes with large time
constants and/or long time delays. If large and frequent
disturbances occur, the process may operate continuously in a
transient state and never attain the desired steady state.
4. In some situations, the controlled variable cannot be
measured on-line, and, consequently, feedback control is not
feasible. 24
Introduction to Feedforward Control
The basic concept of feedforward control is to measure important
disturbance variables and take corrective action before they upset
the process. Feedforward control has several disadvantages:
1. The disturbance variables must be measured on-line. In many
Chapter 15

applications, this is not feasible.


2. To make effective use of feedforward control, at least a crude
process model should be available. In particular, we need to
know how the controlled variable responds to changes in both
the disturbance and manipulated variables. The quality of
feedforward control depends on the accuracy of the process
model.
3. Ideal feedforward controllers that are theoretically capable of
achieving perfect control may not be physically realizable.
Fortunately, practical approximations of these ideal controllers
often provide very effective control.
25
Chapter 15

Figure 15.2 The feedback control of the liquid level in a boiler


drum.
26
A boiler drum with a conventional feedback control system is
shown in Fig. 15.2. The level of the boiling liquid is measured
and used to adjust the feedwater flow rate.
This control system tends to be quite sensitive to rapid changes
in the disturbance variable, steam flow rate, as a result of the
Chapter 15

small liquid capacity of the boiler drum.


Rapid disturbance changes can occur as a result of steam
demands made by downstream processing units.

The feedforward control scheme in Fig. 15.3 can provide better


control of the liquid level. Here the steam flow rate is
measured, and the feedforward controller adjusts the feedwater
flow rate.

27
Chapter 15

Figure 15.3 The feedforward control of the liquid level in a


boiler drum.
28
In practical applications, feedforward control is normally used
in combination with feedback control.
Feedforward control is used to reduce the effects of measurable
disturbances, while feedback trim compensates for inaccuracies
in the process model, measurement error, and unmeasured
Chapter 15

disturbances.

29
Chapter 15

Figure 15.4 The feedfoward-feedback control of the boiler


drum level.
30
Ratio Control
Ratio control is a special type of feedforward control that has
had widespread application in the process industries.
The objective is to maintain the ratio of two process variables
Chapter 15

as a specified value.
The two variables are usually flow rates, a manipulated
variable u, and a disturbance variable d.
Thus, the ratio
u
R= (15-1)
d

is controlled rather than the individual variables. In Eq. 15-1,


u and d are physical variables, not deviation variables.

31
Typical applications of ratio control:
1. Setting the relative amounts of components in blending
operations
2. Maintaining a stoichiometric ratio of reactants to a reactor
Chapter 15

3. Keeping a specified reflux ratio for a distillation column


4. Holding the fuel-air ratio to a furnace at the optimum value.

32
Chapter 15

Figure 15.5 Ratio control, Method I.

33
The main advantage of Method I is that the actual ratio R is
calculated.
A key disadvantage is that a divider element must be included
in the loop, and this element makes the process gain vary in a
nonlinear fashion. From Eq. 15-1, the process gain
Chapter 15

R 1
Kp = = (15-2)
u d d

is inversely related to the disturbance flow rate d.


Because of this significant disadvantage, the preferred scheme
for implementing ratio control is Method II, which is shown in
Fig. 15.6.

34
Chapter 15

Figure 15.6 Ratio control, Method II


35
Regardless of how ratio control is implemented, the process
variables must be scaled appropriately.
For example, in Method II the gain setting for the ratio station
Kd must take into account the spans of the two flow
transmitters.
Chapter 15

Thus, the correct gain for the ratio station is


Sd
K R = Rd (15-3)
Su
where Rd is the desired ratio, Su and Sd are the spans of the
flow transmitters for the manipulated and disturbance streams,
respectively.

36
Example 15.1
A ratio control scheme is to be used to maintain a stoichoimetric
ratio of H2 and N2 as the feed to an ammonia synthesis reactor.
Individual flow controllers will be used for both the H2 and N2
streams. Using the information given below, do the following:
Chapter 15

a) Draw a schematic diagram for the ratio control scheme.


b) Specify the appropriate gain for the ratio station, KR.

37
Available Information
i. The electronic flow transmitters have built-in square root
extractors. The spans of the flow transmitters are 30 L/min for
H2 and 15 L/min for N2.
ii. The control valves have pneumatic actuators.
Chapter 15

iii. Each required current-to-pressure (I/P) transducer has a gain


of 0.75 psi/mA.
iv. The ratio station is an electronic instrument with 4-20 mA
input and output signals.

Solution
The stoichiometric equation for the ammonia synthesis reaction is

3H 2 + N 2 2NH3

38
In order to introduce the feed mixture in stoichiometric
proportions, the ratio of the molar flow rates (H2/N2) should be
3:1. For the sake of simplicity, we assume that the ratio of the
molar flow rates is equal to the ratio of the volumetric flow rates.
But in general, the volumetric flow rates also depend on the
temperature and pressure of each stream (cf., the ideal gas law).
Chapter 15

a) The schematic diagram for the ammonia synthesis reaction is


shown in Fig. 15.7. The H2 flow rate is considered to be the
disturbance variable, although this choice is arbitary because
both the H2 and N2 flow rates are controlled. Note that the ratio
station is merely a device with an adjustable gain. The input
signal to the ratio station is dm, the measured H2 flow rate. Its
output signal usp serves as the set point for the N2 flow control
loop. It is calculated as usp = KRdm.

39
Chapter 15

Figure 15.7 Ratio control scheme for an ammonia synthesis


reactor of Example 15.1
40
b) From the stoichiometric equation, it follows that the desired
ratio is Rd = u/d = 1/3. Substitution into Equation 15-3 gives:

1 30 L / min 2
K R = =
3 15 L / min 3
Chapter 15

Feedforward Controller Design Based on


Steady-State Models
A useful interpretation of feedforward control is that it
continually attempts to balance the material or energy that must
be delivered to the process against the demands of the load.
For example, the level control system in Fig. 15.3 adjusts the
feedwater flow so that it balances the steam demand.
Thus, it is natural to base the feedforward control calculations
on material and energy balances.
41
Chapter 15

Figure 15.8 A simple schematic diagram of a distillation


column.
42
To illustrate the design procedure, consider the distillation
column shown in Fig. 15.8 which is used to separate a binary
mixture.
In Fig. 15.8, the symbols B, D, and F denote molar flow rates,
whereas x, y, and z are the mole fractions of the more volatile
component.
Chapter 15

The objective is to control the distillation composition, y,


despite measurable disturbances in feed flow rate F and feed
composition z, by adjusting distillate flow rate, D.
It is assumed that measurements of x and y are not available.

The steady-state mass balances for the distillation column can be


written as
F = D+B (15-4)
Fz = Dy + Bx (15-5)
43
Solving (15-4) for D and substituting into (15-5) gives
F ( z x)
D= (15-6)
yx
Because x and y are not measured, we replace these variables by
Chapter 15

their set points to yield the feedforward control law:

F ( z xsp )
D= (15-7)
ysp xsp

44
Blending System
Consider the blending system and feedforward controller shown
in Fig. 15.9.
We wish to design a feedforward control scheme to maintain
Chapter 15

exit composition x at a constant set point xsp, despite


disturbances in inlet composition, x1.
Suppose that inlet flow rate w1 and the composition of the other
inlet stream, x2, are constant.
It is assumed that x1 is measured but x is not.

45
Chapter 15

Figure 15.9 Feedforward control of exit composition in the


blending system.
46
The starting point for the feedforward controller design is the
steady-state mass and component balances,
w = w1 + w2 (15-8)
w x = w1 x1 + w2 x2 (15-9)
where the bar over the variable denotes a steady-state value.
Chapter 15

Substituting Eq. 15-8 into 15-9 and solving for w2 gives:


w1 ( x x1 )
w2 = (15-10)
x2 x
In order to derive a feedforward control law, we replace x by xsp,
and w2 and x1 , by w2(t) and x1(t), respectively:
w1 xsp x1 (t )
w2 (t ) = (15-11)
x2 xsp
Note that this feedforward control law is based on the physical
variables rather than on the deviation variables.
47
The feedforward control law in Eq. 15-11 is not in the final
form required for actual implementation because it ignores
two important instrumentation considerations:
First, the actual value of x1 is not available but its measured
value, x1m, is.
Chapter 15

Second, the controller output signal is p rather than inlet flow


rate, w2.
Thus, the feedforward control law should be expressed in
terms of x1m and p, rather than x1 and w2.
Consequently, a more realistic feedforward control law should
incorporate the appropriate steady-state instrument relations
for the w2 flow transmitter and the control valve.

48
Chapter 15
49

Vous aimerez peut-être aussi