30079_28a.pdf

(1491 KB) Pobierz
CHAPTER 28
BASIC CONTROL SYSTEMS DESIGN
William J. Palm III
Mechanical Engineering Department
University of Rhode Island
Kingston, Rhode Island
28.1 INTRODUCTION
868
28.7.3 The Ziegler-Nichols
Rules
891
28.2 CONTROL SYSTEM
STRUCTURE
28.7.4 Nonlinearities and
Controller Performance 892
28.7.5 Reset Windup
869
28.2.1 A Standard Diagram
870
893
28.2.2 Transfer Functions
870
28.2.3 System-Type Number and
Error Coefficients
28.8 COMPENSATION AND
ALTERNATIVE CONTROL
STRUCTURES 893
28.8.1 Series Compensation 893
28.8.2 Feedback
Compensation
and Cascade Control 893
28.8.3 Feedforward
Compensation 894
28.8.4 State- Variable Feedback 895
28.8.5 Pseudoderivative
Feedback
871
28.3 TRANSDUCERS AND ERROR
DETECTORS
872
28.3.1 Displacement and Velocity
Transducers 872
28.3.2 Temperature Transducers 874
28.3.3 Flow Transducers
874
28.3.4 Error Detectors
874
28.3.5 Dynamic Response of
Sensors
875
896
28.4 ACTUATORS
875
28.9 GRAPHICAL DESIGN
METHODS
28.4.1 Electromechanical
Actuators 875
28.4.2 Hydraulic Actuators 876
28.4.3 Pneumatic Actuators 878
896
28.9.1 The Nyquist Stability
Theorem
896
28.9.2 Systems with Dead-Time
Elements
898
28.5 CONTROL LAWS
880
28.9.3 Open-Loop Design for
PID Control
28.5.1 Proportional Control
881
898
28.5.2 Integral Control
883
28.9.4 Design with the Root
Locus
28.5.3 Proportional-Plus-Integral
Control
899
884
28.5.4 Derivative Control
884
28.10 PRINCIPLES OF DIGITAL
CONTROL
28.5.5 PID Control
885
901
28.10.1 Digital Controller
Structure
28.6 CONTROLLER HARDWARE 886
28.6.1 Feedback Compensation
and Controller Design 886
28.6.2 Electronic Controllers 886
28.6.3 Pneumatic Controllers 887
28.6.4 Hydraulic Controllers 887
902
28.10.2 Digital Forms of PID
Control
902
28.11 UNIQUELY DIGITAL
ALGORITHMS
903
28. 1 1 . 1 Digital Feedforward
Compensation
28.7 FURTHER CRITERIA FOR
GAIN SELECTION 887
28.7.1 Performance Indices 889
28.7.2 Optimal Control Methods 891
904
28.11.2 Control Design in the
z-Plane
904
28. 1 1 .3 Direct Design of Digital
Algorithms
908
Revised from William J. Palm III, Modeling, Analysis and Control of Dynamic Systems, Wiley, 1983,
by permission of the publisher.
Mechanical Engineers'Handbook, 2nd ed., Edited by Myer Kutz.
ISBN 0-471-13007-9 © 1998 John Wiley & Sons, Inc.
867
815046027.007.png 815046027.008.png 815046027.009.png 815046027.010.png 815046027.001.png
 
28.12 HARDWARE AND
SOFTWARE FOR DIGITAL
CONTROL
28.13 FUTURE TRENDS IN
CONTROL SYSTEMS 912
28.13.1 Fuzzy Logic Control 913
28.13.2 Neural Networks
909
28.12.1 Digital Control
Hardware
914
909
28.13.3 Nonlinear Control
914
28.12.2 Software for Digital
Control
28.13.4 Adaptive Control
914
911
28.13.5 Optimal Control
914
28.1 INTRODUCTION
The purpose of a control system is to produce a desired output. This output is usually specified by
the command input, and is often a function of time. For simple applications in well-structured situ-
ations, sequencing devices like timers can be used as the control system. But most systems are not
that easy to control, and the controller must have the capability of reacting to disturbances, changes
in its environment, and new input commands. The key element that allows a control system to do
this is feedback, which is the process by which a system's output is used to influence its behavior.
Feedback in the form of the room-temperature measurement is used to control the furnace in a
thermostatically controlled heating system. Figure 28.1 shows the feedback loop in the system's block
diagram, which is a graphical representation of the system's control structure and logic. Another
commonly found control system is the pressure regulator shown in Fig. 28.2.
Feedback has several useful properties. A system whose individual elements are nonlinear can
often be modeled as a linear one over a wider range of its variables with the proper use of feedback.
This is because feedback tends to keep the system near its reference operation condition. Systems
that can maintain the output near its desired value despite changes in the environment are said to
have good disturbance rejection. Often we do not have accurate values for some system parameter,
or these values might change with age. Feedback can be used to minimize the effects of parameter
changes and uncertainties. A system that has both good disturbance rejection and low sensitivity to
parameter variation is robust. The application that resulted in the general understanding of the prop-
erties of feedback is shown in Fig. 28.3. The electronic amplifier gain A is large, but we are uncertain
of its exact value. We use the resistors Rl and R2 to create a feedback loop around the amplifier, and
pick Rl and R2 to create a feedback loop around the amplifier, and pick Rl and R2 so that AR2/Rl
» 1. Then the input-output relation becomes e0 « R^e^R^^ which is independent of A as long as
A remains large. If Rl and R2 are known accurately, then the system gain is now reliable.
Figure 28.4 shows the block diagram of a closed-loop system, which is a system with feedback.
An open-loop system, such as a timer, has no feedback. Figure 28.4 serves as a focus for outlining
the prerequisites for this chapter. The reader should be familiar with the transfer-function concept
based on the Laplace transform, the pulse-transfer function based on the z-transform, for digital
control, and the differential equation modeling techniques needed to obtain them. It is also necessary
to understand block-diagram algebra, characteristic roots, the final-value theorem, and their use in
evaluating system response for common inputs like the step function. Also required are stability
analysis techniques such as the Routh criterion, and transient performance specifications, such as the
damping ratio £, natural frequency a)n, dominant time constant r, maximum overshoot, settling time,
and bandwidth. The above material is reviewed in the previous chapter. Treatment in depth is given
in Refs. 1, 2, and 3.
Fig. 28.1 Block diagram of the thermostat system for temperature control.1
815046027.002.png 815046027.003.png
Fig. 28.2 Pressure regulator: (a) cutaway view; (b) block diagram.1
28.2 CONTROL SYSTEM STRUCTURE
The electromechanical position control system shown in Fig. 28.5 illustrates the structure of a typical
control system. A load with an inertia / is to be positioned at some desired angle 6r. A dc motor is
provided for this purpose. The system contains viscous damping, and a disturbance torque Td acts
on the load, in addition to the motor torque T. Because of the disturbance, the angular position 6 of
the load will not necessarily equal the desired value 6r. For this reason, a potentiometer, or some
other sensor such as an encoder, is used to measure the displacement 6. The potentiometer voltage
representing the controlled position 0 is compared to the voltage generated by the command poten-
tiometer. This device enables the operator to dial in the desired angle dr. The amplifier sees the
difference e between the two potentiometer voltages. The basic function of the amplifier is to increase
the small error voltage e up to the voltage level required by the motor and to supply enough current
required by the motor to drive the load. In addition, the amplifier may shape the voltage signal in
certain ways to improve the performance of the system.
The control system is seen to provide two basic functions: (1) to respond to a command input
that specifies a new desired value for the controlled variable, and (2) to keep the controlled variable
near the desired value in spite of disturbances. The presence of the feedback loop is vital to both
Fig. 28.3 A closed-loop system.
815046027.004.png
Fig. 28.4 Feedback compensation of an amplifier.
functions. A block diagram of this system is shown in Fig. 28.6. The power supplies required for
the potentiometers and the amplifier are not shown in block diagrams of control system logic because
they do not contribute to the control logic.
28.2.1 A Standard Diagram
The electromechanical positioning system fits the general structure of a control system (Fig. 28.7).
This figure also gives some standard terminology. Not all systems can be forced into this format, but
it serves as a reference for discussion.
The controller is generally thought of as a logic element that compares the command with the
measurement of the output, and decides what should be done. The input and feedback elements are
transducers for converting one type of signal into another type. This allows the error detector directly
to compare two signals of the same type (e.g., two voltages). Not all functions show up as separate
physical elements. The error detector in Fig. 28.5 is simply the input terminals of the amplifier.
The control logic elements produce the control signal, which is sent to the final control elements.
These are the devices that develop enough torque, pressure, heat, and so on to influence the elements
under control. Thus, the final control elements are the "muscle" of the system, while the control
logic elements are the "brain." Here we are primarily concerned with the design of the logic to be
used by this brain.
The object to be controlled is the plant. The manipulated variable is generated by the final control
elements for this purpose. The disturbance input also acts on the plant. This is an input over which
the designer has no influence, and perhaps for which little information is available as to the magnitude,
functional form, or time of occurrence. The disturbance can be a random input, such as wind gust
on a radar antenna, or deterministic, such as Coulomb friction effects. In the latter case, we can
include the friction force in the system model by using a nominal value for the coefficient of friction.
The disturbance input would then be the deviation of the friction force from this estimated value and
would represent the uncertainty in our estimate.
Several control system classifications can be made with reference to Fig. 28.7. A regulator is a
control system in which the controlled variable is to be kept constant in spite of disturbances. The
command input for a regulator is its set point. A follow-up system is supposed to keep the control
variable near a command value that is changing with time. An example of a follow-up system is a
machine tool in which a cutting head must trace a specific path in order to shape the product properly.
This is also an example of a servomechanism, which is a control system whose controlled variable
is a mechanical position, velocity, or acceleration. A thermostat system is not a servomechanism, but
a process-control system, where the controlled variable describes a thermodynamic process. Typically,
such variables are temperature, pressure, flow rate, liquid level, chemical concentration, and so on.
28.2.2 Transfer Functions
A transfer function is defined for each input-output pair of the system. A specific transfer function
is found by setting all other inputs to zero and reducing the block diagram. The primary or command
transfer function for Fig. 28.7 is
Fig. 28.5 Position-control system using a dc motor.1
815046027.005.png
Fig. 28.6 Block diagram of the position-control system shown in Fig. 28.5.1
0£) = A(s)Ga(s)Gm(s)Gp(S)
V(s) 1 + Ga(s)Gm(s)Gp(s)H(S)
' }
The disturbance transfer function is
C(s) = ~Q(s)Gp(s)
D(s) 1 + Ga(s)Gm(s)Gp(s)H(s)
V ' ;
The transfer functions of a given system all have the same denominator.
28.2.3 System-Type Number and Error Coefficients
The error signal in Fig. 28.4 is related to the input as
E(s) = * R(s)
(28.3)
1 + G(s)H(s)
If the final value theorem can be applied, the steady-state error is
Elements
Signals
A(s) Input elements B(s) Feedback signal
Ga(s) Control logic elements C(s) Controlled variable or output
Gm(s) Final control elements D(s) Disturbance input
Gp(s) Plant elements E(s) Error or actuating signal
H(s) Feedback elements F(s) Control signal
Q(s) Disturbance elements M(s) Manipulated variable
R(s) Reference input
V(s) Command input
Fig. 28.7 Terminology and basic structure of a feedback-control system.1
815046027.006.png
Zgłoś jeśli naruszono regulamin